Stabilizing and Modulating Color by Copigmentation: Insights from Theory and ExperimentClick to copy article linkArticle link copied!
- Patrick Trouillas
- Juan C. Sancho-García
- Victor De Freitas
- Johannes Gierschner
- Michal Otyepka
- Olivier Dangles
Abstract
Natural anthocyanin pigments/dyes and phenolic copigments/co-dyes form noncovalent complexes, which stabilize and modulate (in particular blue, violet, and red) colors in flowers, berries, and food products derived from them (including wines, jams, purees, and syrups). This noncovalent association and their electronic and optical implications constitute the copigmentation phenomenon. Over the past decade, experimental and theoretical studies have enabled a molecular understanding of copigmentation. This review revisits this phenomenon to provide a comprehensive description of the nature of binding (the dispersion and electrostatic components of π–π stacking, the hydrophobic effect, and possible hydrogen-bonding between pigment and copigment) and of spectral modifications occurring in copigmentation complexes, in which charge transfer plays an important role. Particular attention is paid to applications of copigmentation in food chemistry.
SPECIAL ISSUE
This article is part of the
1 Introduction
Scheme 1
Scheme a(A) Noncovalent association of a prototypical anthocyanin pigment and a prototypical flavonoid copigment (intermolecular copigmentation). (B) Prototypical acylated derivatives allowing copigmentation between the anthocyanin moiety and two phenolic acids covalently linked (intramolecular copigmentation).
2 Pigments and Copigments
2.1 Pigments
2.1.1 (Proto)typical Pigments
Figure 1
Figure 1. Chemical structures of anthocyanidins (R3 = OH) and anthocyanins (R3 = glucose).
Figure 2
Figure 2. Chemical structures of pyranoanthocyanins (flavylium cation form).
2.1.2 Chemistry of Anthocyanins
Figure 3
Figure 3. Structural transformations of anthocyanins in acidic to neutral solution. Proton loss from C5-OH omitted for simplicity.

Figure 4
Figure 4. Structural transformations of the 3′,4′-dihydroxy-7-O-β-d-glucopyranosyloxyflavylium ion. (A) Fast formation of the quinonoid base. Spectral measurements immediately after addition of pigment to buffer, pH = 2.0 → pH = 6.0. (B) Slow accumulation of trans-chalcone (Ctrans). Spectral measurements on solutions equilibrated overnight, pH = 2.0 → pH = 4.0. Adapted with permission from ref 25. Copyright 2013 Elsevier Ltd.
2.2 Copigmentation
2.2.1 Types of Copigments
Figure 5
Figure 5. Chemical structures of series of copigments.
2.2.2 Key Historical Milestones
Scheme 2
Scheme a(A) Intermolecular copigmentation, (B) self-association, (C) intramolecular copigmentation in acylated anthocyanins, (D) self-association of acylated anthocyanins, (E) intercalation in intermolecular copigmentation, and (F) copigmentation in metal–anthocyanin complexes.
3 Pigment···Copigment Complexes in Nature
3.1 Copigmentation in Blue Flowers
Figure 6
Figure 6. Chemical structures of three nonmetallic pigment···copigment assemblies.
Figure 7
Figure 7. Chemical structures of two polyacylated anthocyanins.
Figure 8
Figure 8. Chemical structures of polyacylated anthocyanins.
Figure 9
Figure 9. Anthocyanin-flavone C- or O-glycoside pigments with a malonate bridge from leaves of (A) Oxalis triangularis and (B) Allium “Blue Perfume”.
Figure 10
Figure 10. Structure of commelinin. Blue: malonylawobanin (MA), yellow and orange: flavocommelin (FC), red: Mg2+(A). A side view of a left-handed stacking of two MA units that coordinate to different Mg ions. (B) A side view of a copigmentation of MA and FC in a right-handed stacking arrangement. (C) A side view of a left-handed stacking of two FC units. (D) A side view of commelinin. (E) A skew view of a copigmentation between MA and FC. (F) A skew view of the self-association of two FC units. Adapted with permission from ref 1. Copyright 2009 The Royal Society of Chemistry.
Figure 11
Figure 11. Chemical structure of the anthocyanin derived from Ceanothus papillosus petals.
Figure 12
Figure 12. (A) Structure of protocyanins from Centaurea cyanus (cornflower). (B) X-ray diffraction crystal structure, showing the two metal ions Fe3+ and Mg2+ in the center of the supramolecular assembly. Adapted with permission from ref 1. Copyright 2009 The Royal Society of Chemistry.
common name | anthocyanin | copigments | metal | stoichiometry/pHv (vacuolar pH) | λmax (nm) | source | ref |
---|---|---|---|---|---|---|---|
stoichiometric metal-pigment-copigment complexes | |||||||
commelinin | delphin 3-(6-p-coumaroylglucoside)-5-(6-malonylglucoside) | 6-glucoside 4′-O-glucosylapigenin | Mg2+ | 6 A: 6F: 2M | 646 | Commelina communis | 56 |
protocyanin | cyanidin-3-O-(6-O-succinylglucoside)-5-O-glucoside | apigenin-7-O-glucuronide-4′-O-(6-O malonylglucoside) | Mg2+; Fe3+ | 6–8 A: 6–8F: 2M | 676 | Centaurea cyanus (corn flower) | 57 |
protodelphin | delphin 3-(6-p-coumaroylglucoside)-5-(6-malonylglucoside) | 7,4′-O-diglucosylapigenin | Mg2+ | 6 A: 6F: 2M | – | Salvia patens | 58 |
cyanosalvianin | delphin 3-(6-p-coumaroylglucoside)-5-(6-acetylmalonylglucoside) | 7,4′-O-diglucosylapigenin | Mg2+ | 6 A: 6F: 2M | – | Salvia uliginosa | 59 |
nemophilin | petunin 3-(6-p-coumaroylglucoside)-5-(6-malonylglucoside) | 7-O-glucoside 4′-(6-malonylglucosyl)apigenin | Mg2+; Fe3+ | 712 | Nemophila menziesii | 1 | |
7,4′-O-diglucosylapigenin | |||||||
nonstoichiometric metal-pigment-copigment complexes | |||||||
delphinidin-3-glucoside | caffeoyl and coumaroyl derivatives of quinic acid | Al3+, 4,5 equiv | pHv = 3.6 ± 0.3 | 586 | Hydrangea macrophylla (blue color) | 60,61 | |
delphinidin-3-glucoside | caffeoyl and coumaroyl derivatives of quinic acid | Al3+, 1,2 equiv | pHv = 3.3 ± 0.2 | 539 | Hydrangea macrophylla (red color) | 60,61 | |
cyanidin-7-O-glucoside-3-O-(2-O-glucosyl)-glucoside | kaempferol derivatives | Mg2+; Fe3+ | pHv = 4.8 | 646 and 606 | Meconopsis grandis | 62,63 | |
malvidin 3-O-glucoside-5-O-(6-O-acetylglucoside) | 3-glucoside and 3-sophoroside of kaempferol and myricetin | – | pHv > 5.5 | Geranium | 64 | ||
malvidin 3-O-(4-O-p-coumaroyl-rhamnosyl-6-O-glucoside)-5-O-glucoside | isovitexin (apigenin-6-C-glucoside) | – | Iris ensata | 65 | |||
Petunidin, delphinidin |
3.2 Copigmentation in Wine
3.3 Copigmentation in Food and Beverages
4 Experimental Techniques to Analyze Copigmentation
4.1 UV–vis Absorption
Figure 13
Figure 13. Copigmentation of malvin (malvidin 3,5-di-O-β-d-glucoside) by 7-O-sulfoquercetin. Reprinted with permission from ref 107. Copyright 2001 John Wiley & Sons.







Figure 14
Figure 14. pH dependence of the copigmentation hyperchromic shift. Spectroscopic monitoring at λMAX of free flavylium. Simulations from eq 2 using the following parameters: CP concentration = 10 mM, pK′h = 2.5, pKa = 4.0, rAHCP+ = 0.8, rACP = 0.9, rA = 0.7. Hydroxycinnamic acids (red line): KAHCP+ = 3 × 102 M–1, KACP = 2 × 102 M–1. Flavones and flavonols (green line): KAHCP+ = 3 × 103 M–1, KACP = 103 M–1.
Figure 15
Figure 15. Distribution diagram of anthocyanin forms in the presence of a copigment. Parameters: pK′h = 2.5, pKa = 4.0, pigment and copigment concentrations = 10 and 30 mM, respectively. (A) Hydroxycinnamic acid, KAHCP = 3 × 102, KACP = 2 × 102 M–1, pK′hCP (pH of a 1:1 mixture of colored and colorless forms) ≈ 3.5. (B) Flavone or flavonol, KAHCP = 3 × 102 M−1, KAHCP = 3 × 103 M–1, pK′hCP ≈ 5.2.
4.2 Other Analytical Tools
5 Stability in Intermolecular Copigmentation
5.1 Structure–Affinity Relationships
K (M–1)b | ΔG0 (kJ/mol)b | ΔH0 (kJ/mol) | ΔS0 (J/mol K) | pH | ref | |
---|---|---|---|---|---|---|
self-association | ||||||
pelargonin | 125 | –12.0 | 1 | 44 | ||
cyanin | 265 | –13.8 | 1 | 44 | ||
peonin | 325 | –14.3 | 1 | 44 | ||
delphin | 310 | –14.2 | 1 | 44 | ||
petunin | 625 | –16.0 | 1 | 44 | ||
malvin | 620 | –15.9 | 1 | 44 | ||
malvin | 900 (22 °C) | –16.9 | 6 | 45 | ||
malvin | 200 (22 °C) | –13.1 | 9.8 | 45 | ||
myrtillin | 1240 | –17.7 | 1 | 128 | ||
oenin | 976 | –17.1 | 1 | 128 | ||
oenin | 820 (27 °C) | –16.6 | 1 | 75 | ||
malvidin 3-O-β-d-(6-O-p-coumaroyl)-glucoside | 1400 (27 °C) | –18.0 | 1 | 75 | ||
petunin | 900 | –16.9 | 1 | 128 | ||
kuromanin | 700 | –16.2 | 1 | 128 | ||
peonin | 661 | –16.1 | 1 | 128 | ||
callistephin | 404 | –14.9 | 1 | 128 | ||
copigmentation | ||||||
flavonols and flavones | ||||||
malvin···rutin | 2750 | –19.6 | –37.2 | –59.0 | 3.5 | 130 |
malvin···3‴-O-sulforutin | 350 | –14.5 | –33.0 | –62.0 | 3.5 | 130 |
malvin···2‴,3‴-O-(2-carboxy-1-methylethylidene) rutin (2 epimers) | 2320 | –19.2 | –16.7 | 8.4 | 3.5 | 130 |
malvin···2″,3″,4″,2‴,3‴,4‴-hexa-O-succinylrutin | 1060 | –17.3 | –17.1 | 0.5 | 3.5 | 130 |
malvin···di-O-succinylrutin (3 regioisomers) | 1820 | –18.6 | –13.6 | 16.7 | 3.5 | 130 |
malvin··· 4‴-O-succinylrutin | 1700 | –18.4 | –7.9 | 35.3 | 3.5 | 130 |
malvin···quercetin 3′-O-β-d-(2,3,4,6-tetra-O-succinyl) glucoside | 1560 | –18.2 | –23.6 | –18.0 | 3.5 | 107 |
malvin···7-O-sulfoquercetin | 14470 | –23.8 | –37.3 | –45.6 | 3.5 | 107 |
malvin···4′,7-di-O-sulfoquercetin | 8940 | –22.6 | –34.7 | –40.0 | 3.5 | 107 |
pelargonin···isoquercitrin | 1050 (20 °C) | –17.2 | 0.25 | 105 | ||
pelargonin···quercitrin | 1740 (20 °C) | –18.5 | 0.25 | 105 | ||
delphinidin···baicalin | 2020 | –18.8 | –35.2 | –67.8 | 3.65 | 104 |
delphinidin···quercetin-5′-sulfonic acid | 3290 | –20.0 | –51.7 | –106.6 | 3.65 | 104 |
delphinidin···rutin | 1150 | –17.4 | –55.2 | –163.7 | 3.65 | 104 |
flavanols | ||||||
oenin···(−)-epicatechin | 260 | –13.8 | –18 | –14.2 | 3.6 | 31 |
malvin···catechin | 185 | –30.4 | –58.6 | 3.5 | 113 | |
malvin···catechin | 215 | –25.0 | –39.2 | 5.5 | 113 | |
oenin··· procyanidin B3 | 330 | –14.4 | –9.2 | 17.2 | 3.6 | 31 |
oenin···(9S,11R) vinylcatechin dimer | 1930 | –18.7 | 3.5 | 110 | ||
oenin···(9R,11S) vinylcatechin dimer | 5420 | –21.3 | 3.5 | 110 | ||
oenin···procyanidin B3 | 350 | –14.5 | 3.5 | 110 | ||
catechin-(4→8)-oenin···(9S,11R) vinylcatechin dimer (10% EtOH) | 311 | –14.2 | –17.8 | –12.0 | 3.5 | 108 |
catechin-(4→ 8)-oenin···(9R,11S) vinylcatechin dimer (10% EtOH) | 516 | –15.5 | –22.7 | –24.2 | 3.5 | 108 |
catechin-(4→ 8)-oenin···procyanidin B3 (10% EtOH) | 121 | –11.9 | –32.5 | –69.2 | 3.5 | 108 |
hydroxycinnamic acids | ||||||
malvin···p-coumaric acid | 129 | –12.0 | –32.1 | –62.8 | 2.5 | 106 |
malvin···4-O-β-D-glucopyranosylcoumaric acid | 111 | –11.7 | –26.0 | –38.0 | 106 | |
malvin···caffeic acid | 243 | –13.6 | –25.6 | –46.0 | 106 | |
malvin···4-O-β-d-glucopyranosylcaffeic acid | 244 | –13.6 | –21.6 | –31.0 | 106 | |
malvin···ferulic acid | 331 | –14.4 | –29.0 | –50.1 | 3.6 | 106 |
malvin···4-O-β-D-glucopyranosylferulic acid | 277 | –13.9 | –33.8 | –71.0 | 106 | |
peonidin···ferulic acid | 155 (20 °C) | –12.5 | 2.58 | 105 | ||
oenin···ferulic acid | 243 (20 °C) | –13.6 | 2.58 | 105 | ||
pelargonin···ferulic acid | 263 (20 °C) | –13.8 | 2.58 | 105 | ||
cyanin···ferulic acid | 210 (20 °C) | –13.3 | 2.58 | 105 | ||
malvin···chlorogenic acid | 220 | –13.4 | –24.1 | –36.1 | 3.5 | 113 |
malvin···chlorogenic acid | 140 | –12.2 | –18.3 | –20.1 | 5.5 | 113 |
oenin··· chlorogenic acid | 390 | –14.8 | –14.2 | 2.04 | 3.6 | 31 |
hydroxybenzoic acids | ||||||
pelargonin···protocatechuic acid | 68 (20 °C) | –10.5 | 0.25 | 105 | ||
pelargonin···gallic acid | 87 (20 °C) | –11.1 | 0.25 | 105 | ||
malvin···protocatechuic acid | 80 (20 °C) | –10.9 | 3.65 | 105 | ||
oenin···p-hydroxybenzoic acid (12% EtOH) | 12.1 | –6.2 | –10.7 | –15.3 | 3.6 | 49 |
oenin···protocatechuic acid (12% EtOH) | 14.6 | –6.7 | –13.5 | –23.4 | 3.6 | 49 |
oenin···gallic acid (12% EtOH) | 21.2 | –7.6 | –16.9 | –31.9 | 3.6 | 49 |
oenin···vanillic acid (12% EtOH) | 35.6 | –8.9 | –19.7 | –37.1 | 3.6 | 49 |
oenin···syringic acid (12% EtOH) | 50.6 | –9.7 | –21.1 | –38.8 | 3.6 | 49 |
purines and aromatic amino acids | ||||||
malvin···caffeine | 125 | –12.0 | –11.7 | 0.9 | 3.5 | 113 |
malvin···caffeine | 180 | –12.9 | –17.6 | –16.0 | 5.5 | 113 |
malvin···Trp | 64 | –10.3 | –6.8 | 11.8 | 3.5 | 113 |
malvin···TrpOMe | 36 | –8.9 | –12.3 | –11.5 | 3.5 | 113 |
All values were obtained in aqueous solution unless otherwise specified (presence of a certain % of EtOH).
Values at 25 °C, unless otherwise specified.
5.2 Environmental Effects
5.3 Self-Association
5.4 Metal–Anthocyanin Complexes
6 Stability in Intramolecular Copigmentation
6.1 Conformational Folding of Acylated Anthocyanins
Figure 16
Figure 16. Morning glory (Pharbitis nil) anthocyanins (inspired from ref 135).
Figure 17
Figure 17. Distribution diagram for a prototypical acylated anthocyanin forms showing the strong influence of noncovalent dimerization. Parameters: pK′h = 3.0 (assuming partial protection against water addition owing to intramolecular copigmentation in monomer), pKa = 4.0, pigment concentration = 10 mM. Tentative values for the dimerization constants: KAH+/AH+ = 3 × 104, KAH+/A = 8 × 104, KA/A = 5 × 104 M–1.
6.2 Improving the Chemical Stability of Anthocyanins by Copigmentation
7 Driving Forces of Copigmentation
8 Conformation and Thermodynamics in Copigmentation from Molecular Dynamics
8.1 Molecular Dynamics Challenges
8.1.1 Exploring Conformational Space

Figure 18
Figure 18. Schematic (1D) potential energy surface (PES) showing many local minima accessible by 1 ns-long MD simulation and the necessity to properly explore the entire PES to reach the real valley via long MD simulation or biased techniques to force random exploration. (The noncovalent complex represented here is just a pictorial representation.)
8.1.2 Biased vs Unbiased Molecular Dynamics
8.1.3 Estimation of ΔGbinding
Scheme 3

8.2 Molecular Picture of Copigmentation
8.2.1 Molecular Dynamics Based Studies: Early Stages
Figure 19
Figure 19. Molecular dynamics snapshot of the 3-OMe-cyanidin···quercetin noncovalent complex (the pigment 3-OMe-cyanidin in red; the copigment quercetin in yellow), showing the space between both partners depleted from water molecules.
Figure 20
Figure 20. π–π Stacking complexes between quercetin (yellow) and vitamin E (green) in lipid bilayer membranes.
8.2.2 Challenges in the Theoretical Description of Copigmentation Complexes
9 Conformation and Thermodynamics in Copigmentation from Quantum Chemistry
9.1 Computing Copigmentation with DFT




Figure 21
Figure 21. Dispersion energy as a function of the interatomic distance for a set of selected atomic pairs.
Figure 22
Figure 22. Sketch of the two main DFT-based corrections discussed: the statically averaged interaction (top) and the dynamically interacting densities (bottom).
9.2 Other Methods to Compute Copigmentation
method | CCSD(T), CCSD[T] | CCSD, SCS-CCSD | MP2.5, MP2.X | MP2, SCS-MP2 |
---|---|---|---|---|
scaling | O(N7) | O(N6) | O(N6) | O(N5) |
9.3 Quantum Chemistry Studies on Copigmentation: Failures and Successes
9.3.1 Conformational Analysis
Figure 23
Figure 23. Five stable conformations of the 3-OMe-cyanidin··· quercetin noncovalent complex (A) side-view and top view showing the parallel-displaced arrangement; (B) side-view showing the intermolecular hydrogen bonding of conformers 2 and 5; (C) top view showing the supramolecular chirality in conformers 3 and 5; and (D) side-view showing the slight bending in conformers 1 and 5.
Figure 24
Figure 24. Potential energy curve along the z-axis defined perpendicular to the planar π-conjugated system, as obtained by three methods of calculation.
9.3.2 Association Energies
conformer | conformer | conformer | conformer | conformer | |
---|---|---|---|---|---|
1 | 2 | 3 | 4 | 5 | |
HF-3c | –15.8 | –18.1 | –17.8 | –15.8 | –20.1 |
B3P86-D2/def2-QZVP | –12.6 | –16.5 | –14.1 | –13.8 | –17.3 |
B3P86-D3(BJ)def2-QZVP | –14.5 | –18.2 | –16.6 | –15.8 | –19.2 |
B3P86-NL/def2-QZVP | –17.6 | –21.7 | –19.9 | –19.2 | –23.4 |
SCS-S66-MP2/def2-QZVP | –17.6a | –22.6a | –20.0a | –19.4a | –23.5a |
COSMO-B3P86-D2/def2-QZVP | –6.9b | –8.6b | –6.8b | –7.5b | –10.6b |
COSMO-B3P86-D3(BJ)/def2-QZVP | –8.8b | –10.3b | –9.3b | –9.5b | –12.5b |
COSMO-B3P86-NL/def2-QZVP | –11.7b | –13.6b | –12.4b | –12.8b | –16.5b |
Values may be considered as reference for gas phase estimates.
Values for the water phase represented by a continuum solvent.
10 Underlying Concepts of Spectral Shifts in Copigmentation
10.1 Charge Transfer and Underlying Theory

Scheme 4
Scheme aGrey arrows indicate lowest electronic transitions in a one-electron picture. (B) Resulting exciton states for rotated and laterally displaced AA pair (qualitative quantum-chemical corrected Kasha picture at π-stacking distance).






10.2 Theoretical Investigations of Spectral Shifts in Copigmentation
Figure 25
Figure 25. MO orbital diagram related to the maximum absorption wavelength HOMO and LUMO (left) and NTO analysis (right). Both schemes show the strong CT character of S0 → S1.
10.3 Environmental Effects and Peak Broadening



11 Concluding Remarks, Future Needs, and Applications
Biographies
Patrick Trouillas
Patrick Trouillas received his Ph.D. degree in solid-state physics (optoelectronic properties of fullerene) at the University of Limoges (France). He then joined CEA for a postdoctoral position dealing with thermal exchanges in thin solid films. In 2000, he reoriented his research toward natural compounds (e.g., polyphenols) and pharmaceutical/cosmetic applications. He is currently Assistant Professor at the University of Limoges, jointly affiliated to (i) INSERM-U850 (Limoges) and (ii) the Regional Centre of Advanced Technologies and Materials in Olomouc (Czech Republic). Within active EU networks, he has extensively worked on Molecular Modeling as a tool to rationalize: (i) drug-lipid bilayer interaction and membrane crossing; (ii) biological (antioxidant) properties of natural compounds; and (iii) pigments and dyes' molecular properties. This gave P.T. the opportunity to meet J-C.S-G, J.G., and M.O., with whom he has developed many research projects; he met V.F. and O.D. within the international polyphenol community.
Juan C. Sancho-García
Juan-Carlos Sancho-Garcı́a completed a Ph.D. in Quantum and Computational Chemistry in 2001. During that period, he also held (1999) a Research Fellowship of the J. Heyrovský Institute of Physical Chemistry in Prague (Czech Republic). He later worked (2002–2004) as a Marie Curie Fellowship at the Laboratory for Chemistry of Novel Materials in Mons (Belgium). He obtained (2005) a tenured-track “Ramón y Cajal” Postdoctoral Senior Grant at the University of Alicante, becoming next (2010) an Associate Professor and Lecturer in Physical Chemistry. He has also recently served as an external scientific advisor for high-tech companies and research hubs, such as the Samsung Advanced Institute of Technology, and led as Principal Investigator research projects uninterruptedly from 2004. His main research interest has always been the development and application of DFT methods to challenging systems, in various fields of Chemistry and Materials Science. He met Patrick Trouillas in 2009 and started to work on pigmentation and copigmentation issues, after some preliminary theoretical studies on polyphenol compounds. This fruitful collaboration crystallized in joint publications and in the associated training of young researchers.
Victor De Freitas
Victor De Freitas graduated in Chemistry from the Faculty of Science of the University of Porto (FCUP) in 1984. In 1995, he obtained his Ph.D. in Biological and Medical Sciences in the University of Bordeaux II (France), with a specialization in Oenology. After his Ph.D., he returned to the Department of Chemistry and Biochemistry (DQB) of FCUP, where he has been developing his teaching and research activities. He is currently Full Professor at the University of Porto and member of the REQUIMTE-LAQV Research Centre, where he has been developing an independent area of research on polyphenol compounds: (i) chemical transformations resulting from oxidation processes of polyphenolic pigments, including technological applications in the food industry; (ii) interaction of different classes of polyphenols with proteins in the sensory and nutritional context; and (iii) antioxidant and biological properties of polyphenolic compounds.
Johannes Gierschner
Johannes Gierschner received his Ph.D. in Physical Chemistry in Tübingen, Germany (2000). After stays in Tübingen, Mons, and Georgia Tech, Atlanta, GA, he joined IMDEA Nanoscience in 2008 as a senior researcher (Ramón y Cajal fellow, 2008–2013). From 2008–2010, he was a visiting researcher at the University of Valencia, and since 2009, he is regular visiting researcher at Seoul National University. In 2013, he completed his habilitation (Priv. Doz., University Tübingen). His work integrates optical spectroscopy and computational chemistry to elucidate structure–property relationships in conjugated organic materials.
Michal Otyepka
Michal Otyepka received his Ph.D. degree in Physical Chemistry at the Palacký University, Olomouc, Czech Republic (2004). Currently, he is a head of the Department of Physical Chemistry and vice-director of the Regional Centre of Advanced Technologies and Materials, both at the Palacký University, Olomouc. His research is mostly focused on modeling of biomacromolecules, 2D materials and their interactions, and chemistry of graphene derivatives.
Olivier Dangles
Olivier Dangles studied chemistry at the Ecole Normale Supérieure of Cachan, France (1981–1985), and he obtained a Ph.D. in organic chemistry at the University of Paris-Orsay (1989). He was an assistant professor at the University of Strasbourg (1989–1992), and then a researcher at the National Center of Scientific Research (CNRS, 1992–1995), where he devoted his research to the chemistry of anthocyanins in relation to the expression of natural colors (group of Prof. R. Brouillard). In 1995, he was appointed a Prof. of chemistry at the University of Lyon. Since 2000, he has been a Prof. of chemistry at the University of Avignon, and he works in a joint research unit of Avignon University and the National Institute for Agricultural Research (INRA). His main research topics deal with the chemistry of polyphenols and carotenoids in relation to their coloring properties and their effects on human health: (i) mechanisms of oxidation and antioxidant activity; (ii) interactions with proteins; and (iii) lipids and metal ions and chemical synthesis of metabolites.
Acknowledgment
P.T. thanks the “Conseil Régional du Limousin” for financial support and CALI (CAlcul en LImousin). Financial support from the Czech Science Foundation (P208/12/G016), the Ministry of Education, Youth and Sports of the Czech Republic (project LO1305), and the Operational Program Education for Competitiveness-European Social Fund (project CZ.1.07/2.3.00/20.0058 of the Ministry of Education, Youth and Sports of the Czech Republic) is also gratefully acknowledged. The work at IMDEA was supported by the Spanish Ministerio de Economı́a y Competitividad (MINECO; project CTQ2014-58801). P.T. thanks F. Di Meo and G. Fabre for fruitful discussions on noncovalent interactions and B. Chantemargue for giving a last minute helping hand.
AH+ | flavylium cation |
A | neutral quinonoid base |
A– | anionic neutral quinonoid base |
B | hemiketal |
BSSE | basis set superposition error |
BJ | Becke-Johnson |
Ccis | cis-chalcone |
Ctrans | trans-chalcone |
CC2 | second-order approximate coupled cluster singles and doubles |
CBS | complete basis set |
CCSD | coupled-cluster singles and doubles |
CCSD(T) | coupled-cluster singles, doubles, and triples |
CD | circular dichroism |
CI | configuration interaction |
CIE | commission international de l’éclairage |
CASPT2 | complete-active-space perturbation theory |
COSMO | conductor-like solvation model |
CP | copigment |
CT | charge transfer |
dDsC | density-dependent energy dispersion energy correction |
DFT | density functional theory |
DFTB | density functional tight binding |
DLPNO | domain-based local pair-natural-orbitals |
DSSC | dye-sensitized solar cells |
ES | excited state |
GB | generalized born |
HCA | hydroxycinnamic acid |
HF | Hartree−Fock |
IEFPCM | integral equation formalism polarizable continuum model |
LR | linear-response |
MAD | mean absolute deviation |
MD | molecular dynamics |
MDQ | MD with quenching |
MetaD | meta-dynamics |
MM | molecular mechanics |
MM-GBSA | molecular mechanics- Poisson–Boltzmann/surface area |
MM-PBSA | molecular mechanics-generalized born/surface area |
MO | molecular orbital |
MP2 | second order Møller–Plesset perturbation theory |
NMR | nuclear magnetic resonance |
NOE | nuclear Overhauser effect |
NTO | natural transition orbital |
PCM | polarizable continuum model |
PB | Poisson–Boltzmann |
QM | quantum mechanics |
REMD | replica exchange MD (REMD) |
RSH | range-separated hybrid functionals |
SA | surface area |
SCS | spin-component-scaled |
SS-PCM | state-specific polarizable continuum model |
SS | state-specific |
TD-DFT | time-dependent density functional theory |
TD-HF | time-dependent Hartree−Fock |
vdW-TS | van der Waals Tkatchenko-Scheffler correction |
VV10 | Vydrov-Voorhis nonlocal correlation functional |
XC | exchange-correlation |
XDM | exchange-hole dipole moment |
ZINDO | Zerner’s intermediate neglect of differential overlap-ZINDO |
References
This article references 362 other publications.
- 1Yoshida, K.; Mori, M.; Kondo, T. Blue Flower Color Development by Anthocyanins: From Chemical Structure to Cell Physiology Nat. Prod. Rep. 2009, 26 (7) 857– 964 DOI: 10.1039/b800165kGoogle ScholarThere is no corresponding record for this reference.
- 2Boulton, R. The Copigmentation of Anthocyanins and Its Role in the Color of Red Wine: A Critical Review Am. J. Enol. Vitic. 2001, 52 (2) 67– 87Google Scholar2The copigmentation of anthocyanins and its role in the color of red wine: A critical reviewBoulton, RogerAmerican Journal of Enology and Viticulture (2001), 52 (2), 67-87CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review with refs. Copigmentation is a soln. phenomenon in which pigments and other noncolored org. components form mol. assocns. or complexes. It generally results in an enhancement in the absorbance and in some cases, a shift in the wavelength of the max. absorbance of the pigment. Copigmentation has not previously been taken into account in traditional wine color measures, in the relationship between color and pigment anal., or in spectrophotometric assays for anthocyanin content. It is now apparent that copigmentation can account for between 30 and 50% of the color in young wines and that it is primarily influenced by the levels of several specific, noncolored phenolic components or cofactors. Copigmentation is of crit. importance in understanding the relationship between grape compn. and wine color, the variation in color and pigment concn. between wines, and in all reactions involving the anthocyanins during wine aging. This review focuses on the importance of the individual pigments and cofactors, the strength of their interactions, and their relative abundance in grapes and wines. A simple math. anal. of the soln. equil. is developed to explain the nonlinear deviation from Beer's law. When solved for typical wines, this function provides ests. of the apparent assocn. const., K, and the apparent molar extinction of the copigmented form, Ec, in natural mixts. These measures allow the fraction of the anthocyanins which is in the copigmented form to be estd. The significance of this phenomenon on pigment extn. and color retention during fermns., on the rate of subsequent pigment polymn., on the possible protection of anthocyanins from oxidn., and in the possible involvement on perceived mouth-feel and astringency of wines are suggested. Aspects of the copigmentation phenomenon that are poorly understood are identified and some research directions are suggested.
- 3Landrum, J. T. Carotenoids: Physical, Chemical, and Biological Functions and Properties; CRC Press: Boca Raton, 2010.Google ScholarThere is no corresponding record for this reference.
- 4Haslam, E. Practical Polyphenolics; Cambridge University Press, 1998.Google ScholarThere is no corresponding record for this reference.
- 5Lee, D. Nature’s Palette: The Science of Plant Color, Reprint ed.; University Of Chicago Press: Chicago, 2010.Google ScholarThere is no corresponding record for this reference.
- 6Willstätter, R.; Everest, A. E. Untersuchungen Über Die Anthocyane. I. Über Den Farbstoff Der Kornblume Justus Liebigs Ann. Chem. 1913, 401 (2) 189– 232 DOI: 10.1002/jlac.19134010205Google ScholarThere is no corresponding record for this reference.
- 7Asen, S.; Stewart, R. N.; Norris, K. H. Anthocyanin, Flavonol Copigments, and pH Responsible for Larkspur Flower Colour Phytochemistry 1975, 14 (12) 2677– 2682 DOI: 10.1016/0031-9422(75)85249-6Google ScholarThere is no corresponding record for this reference.
- 8Castañeda-Ovando, A.; Pacheco-Hernández, M.; de, L.; Páez-Hernández, M. E.; Rodríguez, J. A.; Galán-Vidal, C. A. Chemical Studies of Anthocyanins: A Review Food Chem. 2009, 113 (4) 859– 871 DOI: 10.1016/j.foodchem.2008.09.001Google Scholar8Chemical studies of anthocyanins: A reviewCastaneda-Ovando, Araceli; Pacheco-Hernandez, Ma. de Lourdes; Paez-Hernandez, Ma. Elena; Rodriguez, Jose A.; Galan-Vidal, Carlos AndresFood Chemistry (2009), 113 (4), 859-871CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier B.V.)A review. Anthocyanins are natural colorants which have raised a growing interest due to their extensive range of colors, innocuous and beneficial health effects. Despite the great potential of application that anthocyanins represent for food, pharmaceutical and cosmetic industries, their use has been limited because of their relative instability and low extn. percentages. Currently, most investigations on anthocyanins are focused on solving these problems, as well as their purifn. and identification. In this paper, the most recent advances in the chem. investigation of the anthocyanins are summarized, emphasizing the effects of pH, co-pigmentation, metal ion complexation and antioxidant activity on their stability.
- 9He, F.; Liang, N.-N.; Mu, L.; Pan, Q.-H.; Wang, J.; Reeves, M. J.; Duan, C.-Q. Anthocyanins and Their Variation in Red Wines I. Monomeric Anthocyanins and Their Color Expression Molecules 2012, 17 (12) 1571– 1601 DOI: 10.3390/molecules17021571Google ScholarThere is no corresponding record for this reference.
- 10Andersen, O. M.; Markham, K. R. Flavonoids: Chemistry, Biochemistry and Applications; CRC Press, 2005.Google ScholarThere is no corresponding record for this reference.
- 11Fulcrand, H.; Dueñas, M.; Salas, E.; Cheynier, V. Phenolic Reactions during Winemaking and Aging Am. J. Enol. Vitic. 2006, 57 (3) 289– 297Google Scholar11Phenolic reactions during winemaking and agingFulcrand, Helene; Duenas, Montserrat; Salas, Erika; Cheynier, VeroniqueAmerican Journal of Enology and Viticulture (2006), 57 (3), 289-297CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review. The reactivity of polyphenols is due to the position of the hydroxyl groups on their arom. nuclei. Ortho-hydroxyl groups promote oxidn. while meta-hydroxyl groups induce electrophilic arom. substitution. Both hydroxylation patterns are encountered in flavonoid structures, on the B and A rings, resp. In addn. to oxidn. and electrophilic arom. substitution, flavonoids undergo nucleophilic addn. on the central C ring when it is pos. charged. Reactions of the A and C rings are pH-dependent. The A ring of flavonoids undergoes a polycondensation reaction mediated by an aldehyde. The products are anthocyanin and flavanol polymers and copolymers constituted of both. Flavanol polymers are not stable and rearrange into vinyl flavanols and xanthylium pigments. Vinyl flavanols can react with the pos. charged C ring of anthocyanins, yielding pyranoanthocyanins, which can also be formed from components that have a reactive double bond, such as carbonyl and ethylene bonds. The pos. charged C ring primarily undergoes direct reactions. Since the pos. charge on the C ring of anthocyanins and flavanols is pH-dependent, their dehydration and interflavan bond cleavage reactions are also pH-dependent. This leads to flavanol-anthocyanin (F-A+) adducts at lower pH values and anthocyanin-flavanol (A+-F) adducts above pH 3.8. Temp. seems to favor formation of the latter.
- 12Cheynier, V.; Dueñas-Paton, M.; Salas, E.; Maury, C.; Souquet, J.-M.; Sarni-Manchado, P.; Fulcrand, H. Structure and Properties of Wine Pigments and Tannins Am. J. Enol. Vitic. 2006, 57 (3) 298– 305Google Scholar12Structure and properties of wine pigments and tanninsCheynier, Veronique; Duenas-Paton, Montserrat; Salas, Erika; Maury, Chantal; Souquet, Jean-Marc; Sarni-Manchado, Pascale; Fulcrand, HeleneAmerican Journal of Enology and Viticulture (2006), 57 (3), 298-305CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review. Grape phenolics are structurally diverse, from simple mols. to oligomers and polymers that are usually designated "tannins," referring to their ability to interact with proteins. Anthocyanin pigments and tannins are particularly important for red wine quality. Their extn. depends on their location in the berry and their soly. All phenolic compds. are unstable and undergo numerous enzymic and chem. reactions. Color and taste changes during red wine aging have been ascribed to anthocyanin-tannin reactions. The structures and properties of tannins and pigmented tannins from these reactions are often misunderstood. Current research on wine phenolic compn. is reviewed, with emphasis on the following issues: (1) reactions of tannins yield both larger polymers and smaller species; (2) anthocyanin reactions can generate colorless species as well as polymeric and small various pigments; (3) some polymeric pigments undergo sulfite bleaching while some low mol. wt. pigments do not; (4) polymers are both sol. and astringent, so the astringency loss during aging may involve cleavage rather than polymn.; and (5) sensory properties of anthocyanins and tannins are modulated by interactions with other wine components.
- 13de Freitas, V.; Mateus, N. Formation of Pyranoanthocyanins in Red Wines: A New and Diverse Class of Anthocyanin Derivatives Anal. Bioanal. Chem. 2011, 401 (5) 1463– 1473 DOI: 10.1007/s00216-010-4479-9Google Scholar13Formation of pyranoanthocyanins in red wines: a new and diverse class of anthocyanin derivativesde Freitas Victor; Mateus NunoAnalytical and bioanalytical chemistry (2011), 401 (5), 1463-73 ISSN:.Pyranoanthocyanins constitute one of the most important classes of anthocyanin-derived pigments occurring naturally in red wine. Nonetheless, correct assignment of their structures and pathways of formation in red wine has been relatively recent--less than two decades. Study of these newly discovered pigments is progressively unfolding the chemical pathways that drive the evolution of red wine colour during ageing. The objective of this paper is to review current knowledge regarding the pathway of formation in red wine of a great variety of pyranoanthocyanin structures, namely carboxypyranoanthocyanins, methylpyranoanthocyanins, pyranoanthocyanin-flavanols, pyranoanthocyanin-phenols, portisins, oxovitisins, and pyranoanthocyanin dimers. The chromatic features of some of the compounds, for example their colour expression and acid-base equilibria in aqueous media, are also discussed.
- 14Mateus, N.; Silva, A. M. S.; Santos-Buelga, C.; Rivas-Gonzalo, J. C.; de Freitas, V. Identification of Anthocyanin-Flavanol Pigments in Red Wines by NMR and Mass Spectrometry J. Agric. Food Chem. 2002, 50 (7) 2110– 2116 DOI: 10.1021/jf0111561Google ScholarThere is no corresponding record for this reference.
- 15Mateus, N.; Oliveira, J.; Haettich-Motta, M.; de Freitas, V. New Family of Bluish Pyranoanthocyanins J. Biomed. Biotechnol. 2004, 2004 (5) 299– 305 DOI: 10.1155/S1110724304404033Google ScholarThere is no corresponding record for this reference.
- 16He, J.; Carvalho, A. R. F.; Mateus, N.; De Freitas, V. Spectral Features and Stability of Oligomeric Pyranoanthocyanin-Flavanol Pigments Isolated from Red Wines J. Agric. Food Chem. 2010, 58 (16) 9249– 9258 DOI: 10.1021/jf102085eGoogle ScholarThere is no corresponding record for this reference.
- 17He, J.; Oliveira, J.; Silva, A. M. S.; Mateus, N.; De Freitas, V. Oxovitisins: A New Class of Neutral Pyranone-Anthocyanin Derivatives in Red Wines J. Agric. Food Chem. 2010, 58 (15) 8814– 8819 DOI: 10.1021/jf101408qGoogle ScholarThere is no corresponding record for this reference.
- 18Chassaing, S.; Isorez, G.; Kueny-Stotz, M.; Brouillard, R. En Route to Color-Stable Pyranoflavylium Pigments—a Systematic Study of the Reaction between 5-Hydroxy-4-Methylflavylium Salts and Aldehydes Tetrahedron Lett. 2008, 49 (49) 6999– 7004 DOI: 10.1016/j.tetlet.2008.09.139Google ScholarThere is no corresponding record for this reference.
- 19Oliveira, J.; Mateus, N.; Freitas, V. de. Synthesis of a New Bluish Pigment from the Reaction of a Methylpyranoanthocyanin with Sinapaldehyde Tetrahedron Lett. 2011, 52 (16) 1996– 2000 DOI: 10.1016/j.tetlet.2011.02.079Google ScholarThere is no corresponding record for this reference.
- 20Brouillard, R.; Dubois, J.-E. Mechanism of the Structural Transformations of Anthocyanins in Acidic Media J. Am. Chem. Soc. 1977, 99 (5) 1359– 1364 DOI: 10.1021/ja00447a012Google Scholar20Mechanism of the structural transformations of anthocyanins in acidic mediaBrouillard, Raymond; Dubois, Jacques-EmileJournal of the American Chemical Society (1977), 99 (5), 1359-64CODEN: JACSAT; ISSN:0002-7863.In acidic aq. media (pH 1-6), there are three forms of malvin: the flavylium cation I (R = β-D-glucopyranosyl), the carbinol, and the quinonoidal base. Equilibrium between the two neutral forms occurs via the flavylium cation and the equil. const. [carbinol]/[base] is 1.6 (± 0.5) × 102 at 4° ;a relaxation technique was used to det. the rate consts for neutralization (endotherm) of the base, for hydration ofI via formation of a C-O bond and proton transfer,and for the reverse of the reactions.
- 21McClelland, R. A.; Gedge, S. Hydration of the Flavylium Ion J. Am. Chem. Soc. 1980, 102 (18) 5838– 5848 DOI: 10.1021/ja00538a024Google Scholar21Hydration of the flavylium ionMcClelland, Robert A.; Gedge, SherrinJournal of the American Chemical Society (1980), 102 (18), 5838-48CODEN: JACSAT; ISSN:0002-7863.Spectral and kinetic studies were carried out on the transformations undergone in aq. soln. by flavylium ions and I (R = H, Me, MeO). Seven species were identified under various pH conditions, i.e., the flavylium ion, 2 pseudobases, 4-hydroxy adducts and the cis-2-hydroxychalcones and their ionized forms. Equil. between the various species are described and rate and equil. consts. were detd.
- 22Pina, F.; Melo, M. J.; Laia, C. A. T.; Parola, A. J.; Lima, J. C. Chemistry and Applications of Flavylium Compounds: A Handful of Colours Chem. Soc. Rev. 2012, 41 (2) 869– 908 DOI: 10.1039/C1CS15126FGoogle Scholar22Chemistry and applications of flavylium compounds: a handful of coloursPina, Fernando; Melo, Maria J.; Laia, Cesar A. T.; Parola, A. Jorge; Lima, Joao C.Chemical Society Reviews (2012), 41 (2), 869-908CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Flavylium compds. are versatile mols. that comprise anthocyanins, the ubiquitous colorants used by Nature to confer color to most flowers and fruits. They have found a wide range of applications in human technol., from the millenary color paints described by the Roman architect Vitruvius, to their use as food additives, combining color and antioxidant effects, and even as light absorbers in solar cells aiming at a greener solar energy conversion. Their rich complexity derives in part from their ability to switch between a variety of species (flavylium cations, neutral quinoidal bases, hemiketals and chalcones, and neg. charged phenolates) by means of external stimuli, such as pH, temp. and light. This crit. review describes (i) the historical advancements in the understanding of the equil. of their chem. reaction networks; (ii) their thermodn. and kinetics; (iii) the mechanisms underlying their color development, such as co-pigmentation and host-guest interactions; (iv) the photophysics and photochem. that lead to photochromism; and (v) applications in solar cells, models for optical memories, photochromic soft materials such as ionic liqs. and gels, and their properties in solid state materials (274 refs.).
- 23Pina, F. Chemical Applications of Anthocyanins and Related Compounds. A Source of Bioinspiration J. Agric. Food Chem. 2014, 62 (29) 6885– 6897 DOI: 10.1021/jf404869mGoogle Scholar23Chemical Applications of Anthocyanins and Related Compounds. A Source of BioinspirationPina, FernandoJournal of Agricultural and Food Chemistry (2014), 62 (29), 6885-6897CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)A review. Independently of the natural or synthetic origin, flavylium derivs. follow the same network of chem. reactions. Actually, the flavylium cation is stable only at low pH values. Increasing the pH gives rise to the formation of several species: quinoidal bases, hemiketal, cis- and trans-chalcones, and their deprotonated forms. A deep knowledge of the thermodn. and kinetics of these species is an essential tool to practical applications of these compds., in particular, in the domain of food chem. In this work the network of chem. reactions involving flavylium derivs. is presented, and the resp. thermodn. and kinetics are discussed in detail, including the math. expressions and a step-by-step procedure to calc. all of the rate and equil. consts. of the system. Examples of systems possessing a high or low cis-trans isomerization barrier are shown. Recent practical applications of anthocyanins and related compds. illustrate the potential of the flavylium-based family of compds.
- 24Nave, F.; Petrov, V.; Pina, F.; Teixeira, N.; Mateus, N.; de Freitas, V. Thermodynamic and Kinetic Properties of a Red Wine Pigment: Catechin-(4,8)-Malvidin-3- O -Glucoside J. Phys. Chem. B 2010, 114 (42) 13487– 13496 DOI: 10.1021/jp104749fGoogle ScholarThere is no corresponding record for this reference.
- 25Mora-Soumille, N.; Al Bittar, S.; Rosa, M.; Dangles, O. Analogs of Anthocyanins with a 3′,4′-Dihydroxy Substitution: Synthesis and Investigation of Their Acid–base, Hydration, Metal Binding and Hydrogen-Donating Properties in Aqueous Solution Dyes Pigm. 2013, 96 (1) 7– 15 DOI: 10.1016/j.dyepig.2012.07.006Google Scholar25Analogs of anthocyanins with a 3',4'-dihydroxy substitution: Synthesis and investigation of their acid-base, hydration, metal binding and hydrogen-donating properties in aqueous solutionMora-Soumille, Nathalie; Al Bittar, Sheiraz; Rosa, Maxence; Dangles, OlivierDyes and Pigments (2013), 96 (1), 7-15CODEN: DYPIDX; ISSN:0143-7208. (Elsevier Ltd.)Glycosides of hydroxylated flavylium ions are proposed as pertinent analogs of anthocyanins, a major class of polyphenolic plant pigments. Anthocyanins with a 3',4'-dihydroxy substitution on the B-ring (catechol nucleus) are esp. important for their metal chelating and electron-donating (antioxidant) capacities. In this work, an efficient chem. synthesis of 3',4'-dihydroxy-7-O-β-D-glucopyranosyloxyflavylium chloride and its aglycon is reported. Then, the ability of the two pigments to undergo proton transfer (formation of colored quinonoid bases) and add water (formation of a colorless chalcone) is investigated: at equil. the colored quinonoid bases (kinetic products) are present in very minor concns. (<10% of the total pigment concn.) compared to the colorless chalcone (thermodn. product). The glucopyranosyloxyflavylium ion appears significantly less acidic than the aglycon. The thermodn. of the overall sequence of flavylium-chalcone conversion is not affected by the β-D-glucosyl moiety while the kinetics appears slower by a factor ca. 8. Although the glucopyranosyl-oxyflavylium ion and its aglycon display similar affinities for Al3+, the Al3+-glucoside complex is more stable than the Al3+-aglycon complex due to the higher sensitivity of the latter to water addn. and conversion into the corresponding chalcone. Finally, the glucopyranosyl-oxyflavylium ion and its aglycon are compared for their ability to reduce the 1,1-diphenyl-2-picrylhydrazyl radical in a mildly acidic water/MeOH (1:1) mixt. as a first evaluation of their antioxidant activity. Glycosidation at C7-OH results in a lower rate const. of first electron transfer to DPPH and a lower stoichiometry (total no. of 1,1-diphenyl-2-picrylhydrazyl radicals reduced per pigment mol.). Anthocyanins are difficult to ext. from plants in substantial amt. However, the analogs investigated in this work are of easy access by chem. synthesis and express the physico-chem. properties typical of anthocyanins. They can thus be regarded as valuable models for investigating the coloring, metal-binding and antioxidant properties of these important natural pigments.
- 26Yan, Q.; Zhang, L.; Zhang, X.; Liu, X.; Yuan, F.; Hou, Z.; Gao, Y. Stabilization of Grape Skin Anthocyanins by Copigmentation with Enzymatically Modified Isoquercitrin (EMIQ) as a Copigment Food Res. Int. 2013, 50 (2) 603– 609 DOI: 10.1016/j.foodres.2011.04.007Google Scholar26Stabilization of grape skin anthocyanins by copigmentation with enzymatically modified isoquercitrin (EMIQ) as a copigmentYan, Qiuli; Zhang, Linhan; Zhang, Xiaofei; Liu, Xuan; Yuan, Fang; Hou, Zhanqun; Gao, YanxiangFood Research International (2013), 50 (2), 603-609CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)The thermal and light stability of grape skin anthocyanins with enzymically modified isoquercitrin (EMIQ) as a copigment was investigated at different pH levels of 3, 4 and 5. The ratios of anthocyanins to EMIQ were 2:1, 1:1, and 1:2 (wt./wt.), resp., in the thermal expts. at 90 °C, and EMIQ concns. (0.25, 0.5, and 1%, wt./wt.) were evaluated resp. in the light expts. Results revealed that the degrdn. of anthocyanins copigmented with EMIQ followed first-order reaction kinetics. The half life of anthocyanins extended significantly with the increase of EMIQ concn. (p < 0.05), moreover, the color stability increased due to the addn. of EMIQ as the total color difference values ΔE* were smaller for the copigmented anthocyanins. The magnitude of the bathochromic (λmax) shifted to the longest wavelength absorption band with the increasing copigment concn. for all pH levels. Results demonstrated that EMIQ was an effective copigment to stabilize grape skin anthocyanins.
- 27Fanzone, M.; González-Manzano, S.; Pérez-Alonso, J.; Escribano-Bailón, M. T.; Jofré, V.; Assof, M.; Santos-Buelga, C. Evaluation of Dihydroquercetin-3-O-Glucoside from Malbec Grapes as Copigment of Malvidin-3-O-Glucoside Food Chem. 2015, 175, 166– 173 DOI: 10.1016/j.foodchem.2014.11.123Google Scholar27Evaluation of dihydroquercetin-3-O-glucoside from Malbec grapes as copigment of malvidin-3-O-glucosideFanzone, Martin; Gonzalez-Manzano, Susana; Perez-Alonso, Joaquin; Escribano-Bailon, Maria Teresa; Jofre, Viviana; Assof, Mariela; Santos-Buelga, CelestinoFood Chemistry (2015), 175 (), 166-173CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Malbec is a wine grape variety of great phenolic potential characterized for its high levels of anthocyanins and dihydroflavonols. To evaluate the possible implication of dihydroflavonols in the expression of red wine color through reactions of copigmentation or condensation, assays were carried out in wine model systems with different malvidin-3-O-glucoside:dihydroquercetin-3-O-glucoside molar ratios. The addn. of increasing levels of dihydroquercetin-3-O-glucoside to a const. malvidin-3-O-glucoside concn. resulted in a hyperchromic effect assocd. with a darkening of the anthocyanin solns., greater quantity of color and visual satn., perceptible to the human eye. Copigmentation and thermodn. measurements showed that dihydroquercetin-3-O-glucoside can act as an anthocyanin copigment, similar to other usual wine components like flavanols or phenolic acids, although apparently less efficient than flavonols. The high levels of dihydroflavonols existing in Malbec wines in relation to other non-anthocyanin phenolics should make this family of compds. particularly important to explain the color expression in Malbec young red wines.
- 28Baranac, J. M.; Petranovic, N. A.; Dimitric-Markovic, J. M. Spectrophotometric Study of Anthocyan Copigmentation Reactions. 2. Malvin and the Nonglycosidized Flavone Quercetin J. Agric. Food Chem. 1997, 45 (5) 1694– 1697 DOI: 10.1021/jf9606114Google Scholar28Spectrophotometric Study of Anthocyan Copigmentation Reactions. 2. Malvin and the Nonglycosidized Flavone QuercetinBaranac, Jelisaveta M.; Petranovic, Nadezda A.; Dimitric-Markovic, Jasmina M.Journal of Agricultural and Food Chemistry (1997), 45 (5), 1694-1697CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Using UV-vis spectrophotometry we have established that a process of copigmentation takes place between an anthocyan mol., malvin chloride (malvidin 3,5-diglucoside), and a nonglycosidized pentahydroxyflavone, quercetin (3,5,7,3',4'-pentahydroxyflavone). The kinetic and thermodn. parameters, by which the process is characterized, were correlated to the structure, i.e., the nature and position of the substituents in the interacting mols.
- 29Dimitrić-Marković, J. M.; Petranović, N. A.; Baranac, J. M. A Spectrophotometric Study of the Copigmentation of Malvin with Caffeic and Ferulic Acids J. Agric. Food Chem. 2000, 48 (11) 5530– 5536 DOI: 10.1021/jf000038vGoogle ScholarThere is no corresponding record for this reference.
- 30Dimitrić-Marković, J. M.; Baranac, J. M.; Brdaric, T. P. Electronic and Infrared Vibrational Analysis of Cyanidin-Quercetin Copigment Complex Spectrochim. Acta, Part A 2005, 62, 673– 680 DOI: 10.1016/j.saa.2005.02.036Google Scholar30Electronic and infrared vibrational analysis of cyanidin-quercetin copigment complexDimitric Markovic, Jasmina M.; Baranac, Jelisaveta M.; Brdaric, Tanja P.Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy (2005), 62A (1-3), 673-680CODEN: SAMCAS; ISSN:1386-1425. (Elsevier B.V.)Copigment complex formation between cyanidin and quercetin, in aq. buffered solns., was studied by electronic absorption and IR vibrational spectroscopies. It was found that the assocn. of cyanidin with quercetin occurred at pH 3.0 and pH 5.0 including cyanidin flavylium ion and anhydrobase transformation forms, resp. The obtained copigmentation const. values of K = 2726.7 (pH 3.0) and K = 1093.1 (pH 5.0) indicated good assocn. ability of the investigated mols. IR spectra revealed the existence of hydrogen bonds in the copigment complexes structures. The anal. of the deconvoluted IR spectra indicated several types of hydrogen bonds, differently formed: the H-O ··· H bonds with the corresponding bands around 3500 cm-1 and bonds formed via H3O+, oxonium, ion of the mols. with the corresponding bands below 3000 cm-1.
- 31Lambert, S. G.; Asenstorfer, R. E.; Williamson, N. M.; Iland, P. G.; Jones, G. P. Copigmentation between Malvidin-3-Glucoside and Some Wine Constituents and Its Importance to Colour Expression in Red Wine Food Chem. 2011, 125 (1) 106– 115 DOI: 10.1016/j.foodchem.2010.08.045Google Scholar31Copigmentation between malvidin-3-glucoside and some wine constituents and its importance to colour expression in red wineLambert, Stephanie G.; Asenstorfer, Robert E.; Williamson, Natalie M.; Iland, Patrick G.; Jones, Graham P.Food Chemistry (2011), 125 (1), 106-115CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Thermodn. parameters for intermol. copigmentation interactions involving malvidin-3-glucoside were detd. by UV/visible spectroscopy at wine pH (pH 3.6). These included assocn. consts., enthalpy and entropy changes, which were measured for chlorogenic acid, caffeic acid, quercetin, quercetin-3-glucoside, (-)-epicatechin, (+)-catechin, procyanidin dimer and seed tannin. Quercetin produced the strongest copigment (KCP = 2900 ± 1300), while the addn. of glucose at position 3 (quercetin-3-glucoside) reduced its effect by almost 10-fold. Malvidin-3-glucoside self-assocn. (KD = 3300 ± 300 mol-1 l) was thermodynamically favored over intermol. interaction with any of the copigments tested. No color enhancement due to self-assocn. was obsd. for malvidin-3-glucoside-derived pigments that cannot enter hydration reactions. In addn., malvidin-3-(6-O-p-coumaryl)glucoside did not show color enhancement suggesting that the p-coumaryl group prevents self-assocn. The malvidin-3-glucoside CD spectrum was not affected by indicated changes in malvidin-3-glucoside concn. These observations demonstrate that self-assocn. of malvidin-3-glucoside is more important than copigmentation in young red wine.
- 32Malaj, N.; De Simone, B. C.; Quartarolo, A. D.; Russo, N. Spectrophotometric Study of the Copigmentation of Malvidin 3-O-Glucoside with P-Coumaric, Vanillic and Syringic Acids Food Chem. 2013, 141 (4) 3614– 3620 DOI: 10.1016/j.foodchem.2013.06.017Google ScholarThere is no corresponding record for this reference.
- 33Willstätter, R.; Zollinger, E. H. XVI. Über Die Farbstoffe Der Weintraube Und Der Heidelbeere, II Justus Liebigs Ann. Chem. 1917, 412 (2) 195– 216 DOI: 10.1002/jlac.19174120207Google ScholarThere is no corresponding record for this reference.
- 34Robinson, G. M.; Robinson, R. A Survey of Anthocyanins. I Biochem. J. 1931, 25 (5) 1687– 1705 DOI: 10.1042/bj0251687Google ScholarThere is no corresponding record for this reference.
- 35Asen, S.; Stewart, R. N.; Norris, K. H.; Massie, D. R. A Stable Blue Non-Metallic Co-Pigment Complex of Delphanin and C-Glycosylflavones in Prof. Blaauw Iris Phytochemistry 1970, 9 (3) 619– 627 DOI: 10.1016/S0031-9422(00)85702-7Google ScholarThere is no corresponding record for this reference.
- 36Asen, S.; Stewart, R. N.; Norris, K. H. Co-Pigmentation Effect of Quercetin Glycosides on Absorption Characteristics of Cyanidin Glycosides and Color of Red Wing Azalea Phytochemistry 1971, 10 (1) 171– 175 DOI: 10.1016/S0031-9422(00)90266-8Google Scholar36Copigmentation effect of quercetin glycosides on absorption characteristics of cyanidin glycosides and color of red wing azaleaAsen, Sam; Stewart, Robert N.; Norris, Karl H.Phytochemistry (Elsevier) (1971), 10 (1), 171-5CODEN: PYTCAS; ISSN:0031-9422.A quercetin 5-methyl ether and 5 quercetin glycosides were isolated from flowers of Red Wing azalea but were found only in trace amts. in an orange sport of this cultivar. The anthocyanins (cyanidin glycosides) extd. from the orange and red flowers were identical, even though the absorbance spectra of the intact cells differed. The absorbance spectrum of the orange sport was simulated with a 10-3M aq. soln. of cyanidin 3,5-diglucoside at the pH of the tissue, 2.8. The absorbance spectrum of Red Wing was matched with cyanidin 3,5-diglucoside at the same concn. and pH, copigmented with the 3-rhamnoside or galactoside of quercetin.
- 37Asen, S.; Stewart, R. N.; Norris, K. H. Co-Pigmentation of Anthocyanins in Plant Tissues and Its Effect on Color Phytochemistry 1972, 11 (3) 1139– 1144 DOI: 10.1016/S0031-9422(00)88467-8Google Scholar37Copigmentation of anthocyanins in plant tissues and its effect on colorAsen, S.; Stewart, R. N.; Norris, K. H.Phytochemistry (Elsevier) (1972), 11 (3), 1139-44CODEN: PYTCAS; ISSN:0031-9422.Glycosides of the 6 common anthocyanidins all formed copigment complexes with flavonoids and other compds. at pH 2-5. The formation of copigment complexes resulted in a bathochromic shift in the visible λmax. of the anthocyanins and a large increase in extinction at pH 3 and higher. These complexes apparently formed with both the red flavylium salts and the purple anhydro bases. The increase in extinction at pH 3-5 was attributed to the stabilizing effect copigmentation had on the anhydro bases. The degree of copigmentation was a function of the concn. of the anthocyanins and the molar ratio of copigments to anthocyanins. Copigmentation offers an explanation for the infinite color variations that occur in flowers in a pH range where anthocyanins alone are virtually colorless.
- 38Asen, S.; Norris, K. H.; Stewart, R. N. Copigmentation of Aurone and Flavone from Petals of Antirrhinum Majus Phytochemistry 1972, 11 (9) 2739– 2741 DOI: 10.1016/S0031-9422(00)86505-XGoogle ScholarThere is no corresponding record for this reference.
- 39Hoshino, T.; Matsumoto, U.; Goto, T. Self-Association of Some Anthocyanins in Neutral Aqueous Solution Phytochemistry 1981, 20 (8) 1971– 1976 DOI: 10.1016/0031-9422(81)84047-2Google ScholarThere is no corresponding record for this reference.
- 40Hoshino, T.; Matsumoto, U.; Harada, N.; Goto, T. Chiral Exciton Coupled Stacking of Anthocyanins: Interpretation of the Origin of Anomalous CD Induced by Anthocyanin Association Tetrahedron Lett. 1981, 22 (37) 3621– 3624 DOI: 10.1016/S0040-4039(01)81976-6Google ScholarThere is no corresponding record for this reference.
- 41Hoshino, T.; Matsumoto, U.; Goto, T. Evidences of the Self-Association of Anthocyanins I. Circular Dichroism of Cyanin Anhydrobase Tetrahedron Lett. 1980, 21 (18) 1751– 1754 DOI: 10.1016/S0040-4039(00)77827-0Google ScholarThere is no corresponding record for this reference.
- 42Hoshino, T.; Matsumoto, U.; Goto, T.; Harada, N. Evidence for the Self-Association of Anthocyanins IV. PMR Spectroscopic Evidence for the Vertical Stacking of Anthocyanin Molecules Tetrahedron Lett. 1982, 23 (4) 433– 436 DOI: 10.1016/S0040-4039(00)86852-5Google ScholarThere is no corresponding record for this reference.
- 43Hoshino, T.; Goto, T. Effects of pH and Concentration on the Self-Association of Malvin Quinonoidal Base: Electronic and Circular Dichroic Studies Tetrahedron Lett. 1990, 31 (11) 1593– 1596 DOI: 10.1016/0040-4039(90)80025-HGoogle ScholarThere is no corresponding record for this reference.
- 44Hoshino, T. Self-Association of Flavylium Cations of Anthocyanidin 3,5-Diglucosides Studied by Circular Dichroism and 1H NMR Phytochemistry 1992, 31 (2) 647– 653 DOI: 10.1016/0031-9422(92)90053-SGoogle Scholar44Anthocyanin self-aggregates. Part 7. Self-association of flavylium cations of anthocyanidin 3,5-diglucosides studied by circular dichroism and proton NMRHoshino, TsutomuPhytochemistry (1992), 31 (2), 647-53CODEN: PYTCAS; ISSN:0031-9422.Exciton coupling type CD was induced by increasing anthocyanin concn. in strongly acidic media. The higher the concn., the larger the CD magnitude. All anthocyanidin 3,5-diglucosides exhibited neg. signs in the first Cotton effect, which indicates that all the anthocyanins self-assoc. in a left-handed screw manner. Concn. dependence of proton chem. shifts were also obsd. with increase of anthocyanidin 3,5-diglucosides in a strongly acidic medium; upfield shifts of arom. rings by increasing pigment concns. demonstrated that anthocyanidin nuclei stack vertically. Analyses of proton chem. shifts based on the Dimicoli and Helene equation made possible the detn. of the self-assocn. of flavylium cations. The self-assocn. consts. for the most common natural anthocyanidin 3,5-diglucosides are given.
- 45Hoshino, T. An Approximate Estimate of Self-Association Constants and the Self-Stacking Conformation of Malvin Quinonoidal Bases Studied by 1H NMR Phytochemistry 1991, 30 (6) 2049– 2055 DOI: 10.1016/0031-9422(91)85065-8Google ScholarThere is no corresponding record for this reference.
- 46Brouillard, R.; Mazza, G.; Saad, Z.; Albrecht-Gary, A. M.; Cheminat, A. The Co-Pigmentation Reaction of Anthocyanins: A Microprobe for the Structural Study of Aqueous Solutions J. Am. Chem. Soc. 1989, 111 (7) 2604– 2610 DOI: 10.1021/ja00189a039Google Scholar46The co-pigmentation reaction of anthocyanins: a microprobe for the structural study of aqueous solutionsBrouillard, R.; Mazza, G.; Saad, Z.; Albrecht-Gary, A. M.; Cheminat, A.Journal of the American Chemical Society (1989), 111 (7), 2604-10CODEN: JACSAT; ISSN:0002-7863.Visible absorption spectrometry shows that, in acidic aq. solns., chlorogenic acid (I) forms a loose 1:1 complex with malvin chloride (II). The mol. interaction taking place between these two chem. species is characteristic of the copigmentation reaction of anthocyanins. For the first time the mechanism assocd. with this reaction is established. The equation describing the copigment effect is also given. The copigmentation reaction is a very fast process which is extremely influenced by temp. Increasing the temp. or adding MeOH, HCONH2, or NaCl always reduces the copigment effect. In fact, the extent of copigmentation is strictly under the control of the unique mol. structure of liq. water. Finally, the copigmentation phenomenon, which is widespread in higher plants, constitutes a simple, inexpensive, and very sensitive, microprobe for the structural studies of aq. solns.
- 47Figueiredo, P.; Elhabiri, M.; Saito, N.; Brouillard, R. Anthocyanin Intramolecular Interactions. A New Mathematical Approach To Account for the Remarkable Colorant Properties of the Pigments Extracted from Matthiola Incana J. Am. Chem. Soc. 1996, 118 (20) 4788– 4793 DOI: 10.1021/ja9535064Google ScholarThere is no corresponding record for this reference.
- 48Figueiredo, P.; Elhabiri, M.; Toki, K.; Saito, N.; Dangles, O.; Brouillard, R. New Aspects of Anthocyanin Complexation. Intramolecular Copigmentation as a Means for Colour Loss? Phytochemistry 1996, 41 (1) 301– 308 DOI: 10.1016/0031-9422(95)00530-7Google ScholarThere is no corresponding record for this reference.
- 49Zhang, B.; Liu, R.; He, F.; Zhou, P.-P.; Duan, C.-Q. Copigmentation of Malvidin-3-O-Glucoside with Five Hydroxybenzoic Acids in Red Wine Model Solutions: Experimental and Theoretical Investigations Food Chem. 2015, 170, 226– 233 DOI: 10.1016/j.foodchem.2014.08.026Google Scholar49Copigmentation of malvidin-3-O-glucoside with five hydroxybenzoic acids in red wine model solutions: Experimental and theoretical investigationsZhang, Bo; Liu, Rui; He, Fei; Zhou, Pan-Pan; Duan, Chang-QingFood Chemistry (2015), 170 (), 226-233CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)In the present research, the copigmentations of malvidin-3-O-glucoside with five hydroxybenzoic cofactors (p-hydroxybenzoic acid, protocatechuic acid, gallic acid, vanillic acid, and syringic acid) were investigated. The influence of the concn. of these cofactors and the reaction temp. was examd. The equil. const. (K), stoichiometric ratio (n) and the thermodn. parameters (ΔG°, ΔH°, ΔS°) related to the copigmentation were also reported here. Theor. calcns. were performed to identify the relative arrangement between the pigment and cofactors in the copigmentation complexes. Besides, the comparison of the relative binding free energies (ΔΔGbinding) derived from the theor. calcns. and exptl. data were made, and the binding strength of these copigmentation complexes was discussed with the interaction energies (ΔE). AIM anal. was also used to explore the main driving forces contributing to the copigmentation. In the comparison of the five studied cofactors, syringic acid had a stronger copigmentation effect than the other four phenolic acids investigated.
- 50Teixeira, N.; Cruz, L.; Brás, N. F.; Mateus, N.; Ramos, M. J.; de Freitas, V. Structural Features of Copigmentation of Oenin with Different Polyphenol Copigments J. Agric. Food Chem. 2013, 61 (28) 6942– 6948 DOI: 10.1021/jf401174bGoogle Scholar50Structural Features of Copigmentation of Oenin with Different Polyphenol CopigmentsTeixeira, Natercia; Cruz, Luis; Bras, Natercia F.; Mateus, Nuno; Ramos, Maria Joao; de Freitas, VictorJournal of Agricultural and Food Chemistry (2013), 61 (28), 6942-6948CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The copigmentation binding consts. (K) for the interaction of different copigments with oenin (major red wine anthocyanin) were detd. All tests were performed in a 12% ethanol citrate buffer soln. (0.2 M) at pH 3.5, with an ionic strength adjusted to 0.5 M by the addn. of sodium chloride. Over the past years, several copigmentation studies were made and many copigments were tested, but none of them included prodelphinidin B3 or a dimeric-type adduct like oenin-(O)-catechin, probably due to the difficulty in obtaining them. The data yielded from this study allowed concluding that (a) the presence of a pyrogallol group in the B ring of the flavan-3-ol structure slightly increases the copigmentation potential and (b) within all copigments tested oenin-(O)-catechin was revealed to be the best. According to computational studies performed on epicatechin/oenin, epigallocatechin/oenin, procyanidin B3/oenin, and oenin-(O)-catechin/oenin complexes, the ΔGbinding energy of the oenin-(O)-catechin/oenin complex is the most neg. compared to the other copigmentation complexes, hence being more stable and thermodynamically favored. All structural data show that oenin-(O)-catechin and epigallocatechin are closer to the pigment mol., which is in accordance with these two copigments having the highest exptl. copigmentation binding consts. for oenin.
- 51Fossen, T.; Rayyan, S.; Holmberg, M. H.; Nimtz, M.; Andersen, Ø. M. Covalent Anthocyanin–flavone Dimer from Leaves of Oxalis Triangularis Phytochemistry 2007, 68 (5) 652– 662 DOI: 10.1016/j.phytochem.2006.10.030Google ScholarThere is no corresponding record for this reference.
- 52Saito, N.; Nakamura, M.; Shinoda, K.; Murata, N.; Kanazawa, T.; Kato, K.; Toki, K.; Kasai, H.; Honda, T.; Tatsuzawa, F. Covalent Anthocyanin–flavonol Complexes from the Violet-Blue Flowers of Allium “Blue Perfume. Phytochemistry 2012, 80, 99– 108 DOI: 10.1016/j.phytochem.2012.04.011Google ScholarThere is no corresponding record for this reference.
- 53Hayashi, K.; Abe, Y.; Mitsui, S. Blue Anthocyanin from the Flowers of Commelina, the Crystallisation and Some Properties Thereof Proc. Jpn. Acad. 1958, 34 (6) 373– 378Google Scholar53Anthocyanins. XXX. Blue anthocyanin from flowers of Commelina, the crystallization and some properties thereofHayashi, Kozo; Abe, Yukihide; Mitsui, SeijiProceedings of the Japan Academy (1958), 34 (), 373-8CODEN: PJACAW; ISSN:0021-4280.cf. C.A. 50, 16935g. Brilliant blue, prismatic crystals obtained after repeated fractionation and pptn. with water and EtOH were subjected to qual. chem. analysis and behavior. The substance contained an appreciable amt. of metallic elements of which Mg and K were inherent to the pigment mol. It is readily sol. in water but does not retain its color in org. solvents and is neither acid nor base. Paper electrophoresis indicated the existence of essential structural difference between the blue and red forms of anthocyanin. The org. components seemed to be delphinidin, glucose, and p-coumaric acid.
- 54Shibata, K.; Shibata, Y.; Kasiwagi, I. Studies on Anthocyanins: Color Variationin Anthocyanins J. Am. Chem. Soc. 1919, 41 (2) 208– 220 DOI: 10.1021/ja01459a008Google ScholarThere is no corresponding record for this reference.
- 55Bloor, S. J. Blue Flower Colour Derived from Flavonol-Anthocyanin Co-Pigmentation in Ceanothus Papillosus Phytochemistry 1997, 45 (7) 1399– 1405 DOI: 10.1016/S0031-9422(97)00129-5Google ScholarThere is no corresponding record for this reference.
- 56Tamura, H.; Kondo, T.; Goto, T. The Composition of Commelinin, a Highly Associated Metalloanthocaynin Present in the Blue Flower Petals of Commelina Communis Tetrahedron Lett. 1986, 27 (16) 1801– 1804 DOI: 10.1016/S0040-4039(00)84379-8Google ScholarThere is no corresponding record for this reference.
- 57Kondo, T.; Ueda, M.; Isobe, M.; Goto, T. A New Molecular Mechanism of Blue Color Development with Protocyanin, a Supramolecular Pigment from Cornflower, Centaurea Cyanus Tetrahedron Lett. 1998, 39 (45) 8307– 8310 DOI: 10.1016/S0040-4039(98)01858-9Google ScholarThere is no corresponding record for this reference.
- 58Kondo, T.; Oyama, K.; Yoshida, K. Chiral Molecular Recognition on Formation of a Metalloanthocyanin: A Supramolecular Metal Complex Pigment from Blue Flowers of Salvia Patens Angew. Chem., Int. Ed. 2001, 40 (5) 894– 897 DOI: 10.1002/1521-3773(20010302)40:5<894::AID-ANIE894>3.3.CO;2-#Google ScholarThere is no corresponding record for this reference.
- 59Mori, M.; Kondo, T.; Yoshida, K. Cyanosalvianin, a Supramolecular Blue Metalloanthocyanin, from Petals of Salvia Uliginosa Phytochemistry 2008, 69 (18) 3151– 3158 DOI: 10.1016/j.phytochem.2008.03.015Google ScholarThere is no corresponding record for this reference.
- 60Yoshida, K.; Toyama-Kato, Y.; Kameda, K.; Kondo, T. Sepal Color Variation of Hydrangea Macrophylla and Vacuolar pH Measured with a Proton-Selective Microelectrode Plant Cell Physiol. 2003, 44 (3) 262– 268 DOI: 10.1093/pcp/pcg033Google ScholarThere is no corresponding record for this reference.
- 61Kondo, T.; Toyama-Kato, Y.; Yoshida, K. Essential Structure of Co-Pigment for Blue Sepal-Color Development of Hydrangea Tetrahedron Lett. 2005, 46 (39) 6645– 6649 DOI: 10.1016/j.tetlet.2005.07.146Google ScholarThere is no corresponding record for this reference.
- 62Tanaka, M.; Fujimori, T.; Uchida, I.; Yamaguchi, S.; Takeda, K. A Malonylated Anthocyanin and Flavonols in Blue Meconopsis Flowers Phytochemistry 2001, 56 (4) 373– 376 DOI: 10.1016/S0031-9422(00)00357-5Google ScholarThere is no corresponding record for this reference.
- 63Yoshida, K.; Kitahara, S.; Ito, D.; Kondo, T. Ferric Ions Involved in the Flower Color Development of the Himalayan Blue Poppy, Meconopsis Grandis Phytochemistry 2006, 67 (10) 992– 998 DOI: 10.1016/j.phytochem.2006.03.013Google ScholarThere is no corresponding record for this reference.
- 64Markham, K. R.; Mitchell, K. A.; Boase, M. R. Malvidin-3-O-Glucoside-5-O-(6-Acetylglucoside) and Its Colour Manifestation in “Johnson”s Blue’ and Other “Blue” Geraniums Phytochemistry 1997, 45 (2) 417– 423 DOI: 10.1016/S0031-9422(96)00831-XGoogle ScholarThere is no corresponding record for this reference.
- 65Yabuya, T.; Nakamura, M.; Iwashina, T.; Yamaguchi, M.; Takehara, T. Anthocyanin-Flavone Copigmentation in Bluish Purple Flowers of Japanese Garden Iris (Iris Ensata Thunb.) Euphytica 1997, 98 (3) 163– 167 DOI: 10.1023/A:1003152813333Google ScholarThere is no corresponding record for this reference.
- 66Bimpilas, A.; Tsimogiannis, D.; Balta-Brouma, K.; Lymperopoulou, T.; Oreopoulou, V. Evolution of Phenolic Compounds and Metal Content of Wine during Alcoholic Fermentation and Storage Food Chem. 2015, 178, 164– 171 DOI: 10.1016/j.foodchem.2015.01.090Google Scholar66Evolution of phenolic compounds and metal content of wine during alcoholic fermentation and storageBimpilas, Andreas; Tsimogiannis, Dimitrios; Balta-Brouma, Kalliopi; Lymperopoulou, Theopisti; Oreopoulou, VassilikiFood Chemistry (2015), 178 (), 164-171CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Changes in the principal phenolic compds. and metal content during the vinification process and storage under modified atm. (50% N2, 50% CO2) of Merlot and Syrah wines, from grapes cultivated in Greece, have been investigated. Comparing the variation of metals at maceration process, with the variation of monomeric anthocyanins and flavonols, an inverse relationship was noticed, that can be attributed to complexing reactions of polyphenols with particular trace elements. Cu decreased rapidly, whereas a similar behavior that could be expected for Fe and Mn was not confirmed. Differences in the profile of anthocyanins and flavonols in the fresh Merlot and Syrah wines are reported. During 1 yr of storage monomeric anthocyanins declined almost tenfold, probably due to polymn. reactions and copigmentation. Also, a decrease in flavonol glycosides and increase in the resp. aglycons was obsd., attributed to enzymic hydrolysis. The concn. of total phenols and all metals remained practically const.
- 67Hermosín Gutiérrez, I. Influence of Ethanol Content on the Extent of Copigmentation in a Cencibel Young Red Wine J. Agric. Food Chem. 2003, 51 (14) 4079– 4083 DOI: 10.1021/jf021029kGoogle ScholarThere is no corresponding record for this reference.
- 68Monagas, M.; Bartolomé, B. Anthocyanins and Anthocyanin-Derived Compounds. In Wine Chemistry and Biochemistry; Moreno-Arribas, M. V.; Polo, M. C., Eds.; Springer: New York, 2009; pp 439– 462.Google ScholarThere is no corresponding record for this reference.
- 69Rustioni, L.; Bedgood, D. R.; Failla, O.; Prenzler, P. D.; Robards, K. Copigmentation and Anti-Copigmentation in Grape Extracts Studied by Spectrophotometry and Post-Column-Reaction HPLC Food Chem. 2012, 132 (4) 2194– 2201 DOI: 10.1016/j.foodchem.2011.12.058Google ScholarThere is no corresponding record for this reference.
- 70De Rosso, M.; Tonidandel, L.; Larcher, R.; Nicolini, G.; Dalla Vedova, A.; De Marchi, F.; Gardiman, M.; Giust, M.; Flamini, R. Identification of New Flavonols in Hybrid Grapes by Combined Liquid Chromatography–mass Spectrometry Approaches Food Chem. 2014, 163, 244– 251 DOI: 10.1016/j.foodchem.2014.04.110Google ScholarThere is no corresponding record for this reference.
- 71Kelebek, H.; Canbas, A.; Selli, S. HPLC-DAD–MS Analysis of Anthocyanins in Rose Wine Made From Cv. Öküzgözü Grapes, and Effect of Maceration Time on Anthocyanin Content Chromatographia 2007, 66 (3–4) 207– 212 DOI: 10.1365/s10337-007-0277-8Google ScholarThere is no corresponding record for this reference.
- 72Gordillo, B.; Rodríguez-Pulido, F. J.; González-Miret, M. L.; Quijada-Morín, N.; Rivas-Gonzalo, J. C.; García-Estévez, I.; Heredia, F. J.; Escribano-Bailón, M. T. Application of Differential Colorimetry To Evaluate Anthocyanin–Flavonol–Flavanol Ternary Copigmentation Interactions in Model Solutions J. Agric. Food Chem. 2015, 63 (35) 7645– 7653 DOI: 10.1021/acs.jafc.5b00181Google ScholarThere is no corresponding record for this reference.
- 73Somers, T. C.; Vérette, E. Phenolic Composition of Natural Wine Types. In Wine Analysis; Linskens, P. D. H.-F.; Jackson, P. D. J. F., Eds.; Modern Methods of Plant Analysis; Springer: Berlin, 1988; pp 219– 257.Google ScholarThere is no corresponding record for this reference.
- 74González-Manzano, S.; Dueñas, M.; Rivas-Gonzalo, J. C.; Escribano-Bailón, M. T.; Santos-Buelga, C. Studies on the Copigmentation between Anthocyanins and Flavan-3-Ols and Their Influence in the Colour Expression of Red Wine Food Chem. 2009, 114 (2) 649– 656 DOI: 10.1016/j.foodchem.2008.10.002Google Scholar74Studies on the copigmentation between anthocyanins and flavan-3-ols and their influence in the colour expression of red wineGonzalez-Manzano, Susana; Duenas, Montserrat; Rivas-Gonzalo, Julian C.; Escribano-Bailon, M. Teresa; Santos-Buelga, CelestinoFood Chemistry (2009), 114 (2), 649-656CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier B.V.)With the aim of evaluating the importance of the copigmentation process between anthocyanins and flavanols on the color expression of red wine, assays were carried out in wine model systems with mixts. of compds. obtained from two Vitis vinifera grape varieties (Graciano and Tempranillo). Spectrophotometric and chromatic analyses were performed to evaluate the magnitude of the copigmentation and the modifications induced in the color of the solns. Measurement of the changes in the anthocyanin hydration const. (Kh) was also used to det. the strength of the copigmentation process. All the flavanols assayed induced significant changes in the color, perceptible to the human eye, of the wine-like anthocyanin solns. at concns. similar to those that can exist in red wines. The percentage contribution of the copigmentation with flavanols to the color of the anthocyanin solns. was found to range from 2% to 20%. The extent of this effect was related not only to the concn. of flavanols but also to the qual. compn. of the flavanol prepns., as influenced by the part of the grape (either skin or seed) and the variety considered. Divergences were found between the evaluation of the copigmentation based on chromatic parameters in the CIELAB color space and that based on the measurement at visible λ max, as the latter does not consider the integral color changes produced in the visible spectrum. The results obtained confirmed the importance of the qual. phenolic compn., detd. in the wine by the type of grape and winemaking practices, to the prodn. of an effective copigmentation process.
- 75Fernandes, A.; Brás, N. F.; Mateus, N.; Freitas, V. de. A Study of Anthocyanin Self-Association by NMR Spectroscopy New J. Chem. 2015, 39 (4) 2602– 2611 DOI: 10.1039/C4NJ02339KGoogle ScholarThere is no corresponding record for this reference.
- 76Heras-Roger, J.; Pomposo-Medina, M.; Díaz-Romero, C.; Darias-Martín, J. Copigmentation, Colour and Antioxidant Activity of Single-Cultivar Red Wines Eur. Food Res. Technol. 2014, 239 (1) 13– 19 DOI: 10.1007/s00217-014-2185-0Google Scholar76Copigmentation, colour and antioxidant activity of single-cultivar red winesHeras-Roger, J.; Pomposo-Medina, M.; Diaz-Romero, C.; Darias-Martin, J.European Food Research and Technology (2014), 239 (1), 13-19CODEN: EFRTFO; ISSN:1438-2377. (Springer)A hundred and thirty-six single-cultivar red wines of different vintages were collected from several wineries in the Canary Islands in order to study the magnitude of the copigmentation phenomenon and the antioxidant activity. The contribution of free anthocyanins, copigmented anthocyanins and polymeric pigments to the color of wine, as well as the total phenols, the antioxidant activity (2,2-diphenyl-1-picrylhydrazyl, DPPH method) and the chromatic characteristics of the wines were detd. The influence of ageing time and the climatic conditions on these parameters was also studied. The wines made with Merlot, Ruby Cabernet and Syrah cultivars showed the highest parameters of color, and the largest contribution to the copigmented anthocyanins was from the Ruby Cabernet, Listan negro and Syrah cultivars. The copigmented anthocyanins and the free anthocyanins decrease with the age of the wine, and the antioxidant activity of the samples appears to be related to the total phenol content. An influence of the climatic conditions on color parameters has been found. The correlation study between parameters suggests that the parameters b* and L* could be used as suitable indicators of evolution or oxidn. stage of red wines.
- 77García-Marino, M.; Escudero-Gilete, M. L.; Heredia, F. J.; Escribano-Bailón, M. T.; Rivas-Gonzalo, J. C. Color-Copigmentation Study by Tristimulus Colorimetry (CIELAB) in Red Wines Obtained from Tempranillo and Graciano Varieties Food Res. Int. 2013, 51 (1) 123– 131 DOI: 10.1016/j.foodres.2012.11.035Google Scholar77Color-copigmentation study by tristimulus colorimetry (CIELAB) in red wines obtained from Tempranillo and Graciano varietiesGarcia-Marino, Matilde; Escudero-Gilete, Maria Luisa; Heredia, Francisco Jose; Escribano-Bailon, Maria Teresa; Rivas-Gonzalo, Julian CarlosFood Research International (2013), 51 (1), 123-131CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)A study of the changes of copigmentation phenomenon in wines elaborated from different varieties has been undertaken. Colorimetric measurement of Tempranillo (T) and Graciano (G) monovarietal wines, and two 80:20 blend wines: M, (grape blending T and G, co-maceration) and W (wine blending T and G, co-vinification) was performed by spectrophotometry. Significant differences (p < 0.05) were found among the color of the wines. The Graciano cv. afforded somewhat darker and more colorful wines than the other wines. The color difference values, ΔE*ab suggested that co-vinification (W) led to wines being more similar to T than the co-maceration (M). The ΔE*ab[w-c] between untreated wines - whole wines, w - and the wines dild. to eliminate copigmentation - cor. wines, c - was 14.2 CIELAB units in the initial stages of winemaking and 6.7 in the final stages. M had a greater proportion of color due to copigmentation than the monovarietal wines. Evaluation of this parameter confirms the importance of copigmentation process into wine color during the early stages of the vinification. Also, through the full spectrum, quant. data obtained allow a visual interpretation of the changes involved. In addn., with the aging in bottle, M wines had more stable color and more different color than W wines.
- 78Schwarz, M.; Picazo-Bacete, J. J.; Winterhalter, P.; Hermosín-Gutiérrez, I. Effect of Copigments and Grape Cultivar on the Color of Red Wines Fermented after the Addition of Copigments J. Agric. Food Chem. 2005, 53 (21) 8372– 8381 DOI: 10.1021/jf051005oGoogle ScholarThere is no corresponding record for this reference.
- 79Aleixandre-Tudó, J. L.; Álvarez, I.; Lizama, V.; García, M. J.; Aleixandre, J. L.; Du Toit, W. J. Impact of Caffeic Acid Addition on Phenolic Composition of Tempranillo Wines from Different Winemaking Techniques J. Agric. Food Chem. 2013, 61 (49) 11900– 11912 DOI: 10.1021/jf402713dGoogle ScholarThere is no corresponding record for this reference.
- 80Gris, E. F.; Ferreira, E. A.; Falcão, L. D.; Bordignon-Luiz, M. T. Influence of Ferulic Acid on Stability of Anthocyanins from Cabernet Sauvignon Grapes in a Model System and a Yogurt System Int. J. Food Sci. Technol. 2007, 42 (8) 992– 998 DOI: 10.1111/j.1365-2621.2006.01335.xGoogle Scholar80Influence of ferulic acid on stability of anthocyanins from Cabernet Sauvignon grapes in a model system and a yogurt systemGris, Eliana Fortes; Ferreira, Eduardo Antonio; Falcao, Leila Denise; Bordignon-Luiz, Marilde TerezinhaInternational Journal of Food Science and Technology (2007), 42 (8), 992-998CODEN: IJFTEZ; ISSN:0950-5423. (Blackwell Publishing Ltd.)The influence of different factors on the stability of the anthocyanin crude ext. from Cabernet Sauvignon grape skins was investigated. In a model system, the factors evaluated were as follows: temp. 4 ± 1 °C and 29 ± 3 °C, presence and absence of light, pH 3.0 and 4.0 and presence of ferulic acid. The influence of the addn. of ferulic acid to anthocyanins was investigated in a yogurt system stored at 4 ± 1 °C. The results obtained for anthocyanin degrdn. velocity const. and for the half-life time of anthocyanins in a model system and in a yogurt system showed that ferulic acid significantly increased the stability of the anthocyanins crude ext.
- 81Talcott, S. T.; Peele, J. E.; Brenes, C. H. Red Clover Isoflavonoids as Anthocyanin Color Enhancing Agents in Muscadine Wine and Juice Food Res. Int. 2005, 38 (10) 1205– 1212 DOI: 10.1016/j.foodres.2005.05.004Google Scholar81Red clover isoflavonoids as anthocyanin color enhancing agents in muscadine wine and juiceTalcott, Stephen T.; Peele, Janelle E.; Brenes, Carmen H.Food Research International (2005), 38 (10), 1205-1212CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)Isoflavonoid exts. from red clover (Trifolium pratense) leaves were found to enhance overall color and stability of anthocyanin 3,5-diglucosides present in muscadine grape (Vitis rotundifolia) juice and wine through intermol. copigmentation reactions. Predominant isoflavonoids present in red clover included formononetin, biochanin A, and prunetin and were the major polyphenolics identified to influence anthocyanin color and stability. Since red clover isoflavonoids have poor water soly. characteristics, this allowed for removal of extraneous non-isoflavonoid compds. using hot water and subsequent extn. with ethanol. Isoflavonoid soly. was evaluated as a function of ethanol concn. with recoveries up to 57% found in 20% solns. Changes in max. absorbance, total sol. phenolics, isoflavonoids, and anthocyanins were evaluated in muscadine juice and wine following the addn. of isoflavonoid exts. with max. color enhancement found at an anthocyanin to cofactor ratio of 1:8, after which their soly. was prohibitive. Addnl., dried leaves and ethanolic exts. of red clover were added prior to and following fermn. of muscadine wine (11% ethanol) stimulating the natural copigmentation that takes place during red wine fermn. and aging processes. Once fermn. was complete, finished wines were evaluated over a 9-wk storage period at 20 and 37°. Despite low levels of isoflavonoids present, color improvement and anthocyanin stability was obsd. in the wines during storage. Little information is available on copigmentation reactions occurring in actual food systems, yet red clover isoflavonoids proved to be novel and effective color enhancing compds. when used in low concns. in young muscadine wines.
- 82Liu, S.; Fu, Y.; Nian, S. Buffering Colour Fluctuation of Purple Sweet Potato Anthocyanins to Acidity Variation by Surfactants Food Chem. 2014, 162, 16– 21 DOI: 10.1016/j.foodchem.2014.04.029Google Scholar82Buffering color fluctuation of purple sweet potato anthocyanins to acidity variation by surfactantsLiu, Songbai; Fu, Yuanqing; Nian, SiFood Chemistry (2014), 162 (), 16-21CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Anthocyanins are intriguing natural pigments with beneficial bioactivities and their color is extremely susceptible to acidity variation. Minimization of color fluctuation is essential to maintain quality consistency in food industry. A new strategy employing surfactants to mimic encapsulation was attempted with typical anionic, cationic and nonionic surfactants and proved effective although the traditional copigmentation method was inactive. The exceptional color fluctuation buffering effect of anionic surfactants esp. SDS was revealed and then carefully analyzed by colorimetric and spectroscopic methods. The outstanding activity of SDS presumably resulted from effective shielding of anthocyanins from external acidity through strong interaction with the pos. charged flavylium cations owing to its anionic nature. These results suggest SDS is a valuable additive for buffering color fluctuation of anthocyanins. The strategy of surfactant will be useful for buffering color fluctuation of natural colorants.
- 83Kovac, V.; Alonso, E.; Bourzeix, M.; Revilla, E. Effect of Several Enological Practices on the Content of Catechins and Proanthocyanidins of Red Wines J. Agric. Food Chem. 1992, 40 (10) 1953– 1957 DOI: 10.1021/jf00022a045Google ScholarThere is no corresponding record for this reference.
- 84Canals, R.; del Carmen Llaudy, M.; Canals, J. M.; Zamora, F. Influence of the Elimination and Addition of Seeds on the Colour, Phenolic Composition and Astringency of Red Wine Eur. Food Res. Technol. 2008, 226 (5) 1183– 1190 DOI: 10.1007/s00217-007-0650-8Google Scholar84Influence of the elimination and addition of seeds on the colour, phenolic composition and astringency of red wineCanals, Roser; del Carmen Llaudy, Maria; Canals, Joan Miquel; Zamora, FernandoEuropean Food Research and Technology (2008), 226 (5), 1183-1190CODEN: EFRTFO; ISSN:1438-2377. (Springer GmbH)Several winemaking techniques were developed to eliminate seeds to prevent the release of high amts. of very astringent proanthocyanidins, esp. when the grapes are unripe. However, there is no scientific information on the effects of this practice. The aim of this paper is to study how the elimination and addn. of seeds influence the color, phenolic compn. and astringency of red wine. The elimination of around 80% of the seeds led to a significant decrease of color intensity and anthocyanin concn. The addn. of seeds originated wines with a greater concn. of total anthocyanin, but did not significantly affect the wine color. These wines also presented significantly higher levels of proanthocyanidins, a greater proportion of epicatechin-3-gallate, a lower mean d.p., and, in particular, a drastic increase in astringency. Wines obtained with the elimination of seeds, on the other hand, had exactly the opposite characteristics.
- 85Escribano-Bailon, M. T.; Santos-Buelga, C. Anthocyanin Copigmentation - Evaluation, Mechanisms and Implications for the Colour of Red Wines Curr. Org. Chem. 2012, 16 (6) 715– 723 DOI: 10.2174/138527212799957977Google Scholar85Anthocyanin copigmentation - evaluation, mechanisms and implications for the colour of red winesEscribano-Bailon, Maite T.; Santos-Buelga, CelestinoCurrent Organic Chemistry (2012), 16 (6), 715-723CODEN: CORCFE; ISSN:1385-2728. (Bentham Science Publishers Ltd.)A review. Copigmentation is the main color-stabilizing mechanism in plants and in food products of vegetable origin. It is a spontaneous and exothermic process that consists of the stacking of an org. mol., called copigment, on the planar polarizable moieties of the anthocyanin colored forms. Although this phenomenon has long been described, there are some aspects that are still not well understood or controversial like the nature of the interaction pigment to copigment, the way to quantify the extent of the process, its effect on other anthocyanin properties like astringency or reactivity. In this article a review of the most significant advances achieved in the last years in the field of intramol. and intermol. copigmentation is presented. Also, the most recent findings regarding wine copigments and their effects on the color of red wines are revised.
- 86Quijada-Morín, N.; Dangles, O.; Rivas-Gonzalo, J. C.; Escribano-Bailón, M. T. Physico-Chemical and Chromatic Characterization of Malvidin 3-Glucoside-Vinylcatechol and Malvidin 3-Glucoside-Vinylguaiacol Wine Pigments J. Agric. Food Chem. 2010, 58 (17) 9744– 9752 DOI: 10.1021/jf102238vGoogle ScholarThere is no corresponding record for this reference.
- 87Oliveira, J.; Mateus, N.; Silva, A. M. S.; de Freitas, V. Equilibrium Forms of Vitisin B Pigments in an Aqueous System Studied by NMR and Visible Spectroscopy J. Phys. Chem. B 2009, 113 (32) 11352– 11358 DOI: 10.1021/jp904776kGoogle ScholarThere is no corresponding record for this reference.
- 88Brouillard, R.; Dangles, O. Anthocyanin Molecular Interactions: The First Step in the Formation of New Pigments during Wine Aging? Food Chem. 1994, 51 (4) 365– 371 DOI: 10.1016/0308-8146(94)90187-2Google ScholarThere is no corresponding record for this reference.
- 89Wrolstad, R. E.; Erlandson, J. A. Effect of Metal Ions on the Color of Strawberry Puree J. Food Sci. 1973, 38 (3) 460– 463 DOI: 10.1111/j.1365-2621.1973.tb01454.xGoogle ScholarThere is no corresponding record for this reference.
- 90Starr, M. S.; Francis, F. J. Effect of Metallic Ions on Color and Pigment Content of Cranberry Juice Cocktail J. Food Sci. 1973, 38 (6) 1043– 1046 DOI: 10.1111/j.1365-2621.1973.tb02144.xGoogle ScholarThere is no corresponding record for this reference.
- 91Kallio, H.; Pallasaho, S.; Kärppä, J.; Linko, R. R. Comparison of the Half-Lives of the Anthocyanins in the Juice of Crowberry, Empetrum Nigrum J. Food Sci. 1986, 51 (2) 408– 410 DOI: 10.1111/j.1365-2621.1986.tb11142.xGoogle ScholarThere is no corresponding record for this reference.
- 92Osawa, Y. Chapter 2 - Copigmentation of Anthocyanins. In Anthocyanins As Food Colors; Markakis, P., Ed.; Academic Press, 1982; pp 41– 68.Google ScholarThere is no corresponding record for this reference.
- 93Rein, M. J.; Heinonen, M. Stability and Enhancement of Berry Juice Color J. Agric. Food Chem. 2004, 52 (10) 3106– 3114 DOI: 10.1021/jf035507iGoogle ScholarThere is no corresponding record for this reference.
- 94Maccarone, E.; Maccarrone, A.; Rapisarda, P. Stabilization of Anthocyanins of Blood Orange Fruit Juice J. Food Sci. 1985, 50 (4) 901– 904 DOI: 10.1111/j.1365-2621.1985.tb12976.xGoogle ScholarThere is no corresponding record for this reference.
- 95Wilska-Jeszka, J.; Korzuchowska, A. Anthocyanins and Chlorogenic Acid Copigmentation - Influence on the Colour of Strawberry and Chokeberry Juices Z. Lebensm.-Unters. Forsch. 1996, 203 (1) 38– 42 DOI: 10.1007/BF01267767Google Scholar95Anthocyanins and chlorogenic acid copigmentation. Influence on the color of strawberry and chokeberry juicesWilska-Jeszka, Jadwiga; Korzuchowska, AnnaZeitschrift fuer Lebensmittel-Untersuchung und -Forschung (1996), 203 (1), 38-42CODEN: ZLUFAR; ISSN:0044-3026. (Springer)The effect of copigmentation on chlorogenic acid with anthocyanins in strawberry and chokeberry juices was investigated. Chlorogenic acid, at concns. greater than that of anthocyanins, enhanced the color intensity of these juices. The max. copigmentation effect in both juices was obsd. at pH 3.4. In the investigated range of the copigment/pigment ratio, i.e. 1:1 to 50:1, absorbance increased (ΔA) linearly with copigment content, ΔA/g chlorogenic acid was greater in chokeberry than in strawberry juices. In solns. of purified pigments of these fruits, smaller copigmentation effects were obsd. than in juices under the same conditions, which indicated the participation of natural copigments present in fruits in the copigmentation process.
- 96Hernández-Herrero, J. A.; Frutos, M. J. Influence of Rutin and Ascorbic Acid in Colour, Plum Anthocyanins and Antioxidant Capacity Stability in Model Juices Food Chem. 2015, 173, 495– 500 DOI: 10.1016/j.foodchem.2014.10.059Google ScholarThere is no corresponding record for this reference.
- 97Sari, P.; Wijaya, C. H.; Sajuthi, D.; Supratman, U. Colour Properties, Stability, and Free Radical Scavenging Activity of Jambolan (Syzygium Cumini) Fruit Anthocyanins in a Beverage Model System: Natural and Copigmented Anthocyanins Food Chem. 2012, 132 (4) 1908– 1914 DOI: 10.1016/j.foodchem.2011.12.025Google ScholarThere is no corresponding record for this reference.
- 98Fischer, U. A.; Carle, R.; Kammerer, D. R. Thermal Stability of Anthocyanins and Colourless Phenolics in Pomegranate (Punica Granatum L.) Juices and Model Solutions Food Chem. 2013, 138 (2–3) 1800– 1809 DOI: 10.1016/j.foodchem.2012.10.072Google ScholarThere is no corresponding record for this reference.
- 99Pan, Y.-Z.; Guan, Y.; Wei, Z.-F.; Peng, X.; Li, T.-T.; Qi, X.-L.; Zu, Y.-G.; Fu, Y.-J. Flavonoid C-Glycosides from Pigeon Pea Leaves as Color and Anthocyanin Stabilizing Agent in Blueberry Juice Ind. Crops Prod. 2014, 58, 142– 147 DOI: 10.1016/j.indcrop.2014.04.029Google Scholar99Flavonoid C-glycosides from pigeon pea leaves as color and anthocyanin stabilizing agent in blueberry juicePan, You-Zhi; Guan, Yue; Wei, Zuo-Fu; Peng, Xiao; Li, Ting-Ting; Qi, Xiao-Lin; Zu, Yuan-Gang; Fu, Yu-JieIndustrial Crops and Products (2014), 58 (), 142-147CODEN: ICRDEW; ISSN:0926-6690. (Elsevier B.V.)The influences of flavonoid C-glycoside exts. from pigeon pea leaves (FCGE) and its main components vitexin and orientin on the color and anthocyanins stability of blueberry juice were investigated in the thermal expts. The color of juice was enhanced by FCGE and its main components in a molar ratio of 1:1 (anthocyanins/copigment). The satd. color of juice with FCGE and its main components could hold for a significantly longer time. Addnl., copigment FCGE, vitexin and orientin significantly enhanced the stability of anthocyanins. Half-life of anthocyanins in juice samples with FCGE (1:1), orientin and vitexin increased 87%, 79%, and 62%, resp. These results indicated FCGE showed dramatic effect on the color and anthocyanins stability of juice. Furthermore, juice samples with copigments possessed higher total phenolics content and higher DPPH radical-scavenging activity. Therefore, natural, easily available, inexpensive FCGE has potential as color enhancer and anthocyanins stabilizer in food industry.
- 100Goto, T.; Kondo, T. Struktur Und Molekulare Stapelung von Anthocyanen—Variation Der Blütenfarben Angew. Chem. 1991, 103 (1) 17– 33 DOI: 10.1002/ange.19911030105Google ScholarThere is no corresponding record for this reference.
- 101Liao, H.; Cai, Y.; Haslam, E. Polyphenol Interactions. Anthocyanins: Co-Pigmentation and Colour Changes in Red Wines J. Sci. Food Agric. 1992, 59 (3) 299– 305 DOI: 10.1002/jsfa.2740590305Google Scholar101Polyphenol interactions. Part 6. Anthocyanins: co-pigmentation and color changes in red winesLiao, Hua; Cai, Ya; Haslam, EdwinJournal of the Science of Food and Agriculture (1992), 59 (3), 299-305CODEN: JSFAAE; ISSN:0022-5142.Anthocyanin co-pigmentation is reviewed and its relevance to the color and appearance of young red wines is outlined. Reactions between flavan-3-ols, such as are present in red wines, and the anthocyanin malvin to give yellow-orange pigments (absorbance max. ∼440 nm) are reported. Possible structures for these new pigments are discussed as is their probable contribution to the changing hues of red wines as they age.
- 102Brouillard, R.; Wigand, M.-C.; Dangles, O.; Cheminat, A. pH and Solvent Effects on the Copigmentation Reaction of Malvin with Polyphenols, Purine and Pyrimidine Derivatives J. Chem. Soc., Perkin Trans. 2 1991, 8) 1235– 1241 DOI: 10.1039/p29910001235Google Scholar102The pH and solvent effects on the copigmentation reaction of malvin with polyphenols, purine and pyrimidine derivativesBrouillard, Raymond; Wigand, Marie Claude; Dangles, Olivier; Cheminat, AnnieJournal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1991), (8), 1235-41CODEN: JCPKBH; ISSN:0300-9580.The influence of pH on the copigmentation reaction of malvin has been investigated from an exptl. and theor. viewpoint. The general equation for the copigment effect, when monitored by visible absorption spectrometry, is derived and is in good agreement with results obtained in the case of three different copigments, namely chlorogenic acid, caffeine and adenosine. In particular, it is demonstrated that assocn. of malvin with the copigment occurs for all colored malvin species and the corresponding stability consts. are given. Some tannins and purine or pyrimidine derivs. have also been tested for their ability to act as copigments; some assoc. quite strongly with malvin. Water exerted the greatest effect on the extent of the copigmentation phenomenon. Such a result seems to indicate that the strength of the copigment effect parallels the cohesion of the H-bonded tetrahedral network of water mols.
- 103Baranac, J. M.; Petranovic, N. A.; Dimitric-Markovic, J. M. Spectrophotometric Study of Anthocyan Copigmentation Reactions J. Agric. Food Chem. 1996, 44 (5) 1333– 1336 DOI: 10.1021/jf950420lGoogle ScholarThere is no corresponding record for this reference.
- 104Oszmiański, J.; Bakowska, A.; Piacente, S. Thermodynamic Characteristics of Copigmentation Reaction of Acylated Anthocyanin Isolated from Blue Flowers ofScutellaria Baicalensis Georgi with Copigments J. Sci. Food Agric. 2004, 84 (12) 1500– 1506 DOI: 10.1002/jsfa.1815Google Scholar104Thermodynamic characteristics of copigmentation reaction of acylated anthocyanin isolated from blue flowers of Scutellaria baicalensis Georgi with copigmentsOszmianski, Jan; Bakowska, Anna; Piacente, SoniaJournal of the Science of Food and Agriculture (2004), 84 (12), 1500-1506CODEN: JSFAAE; ISSN:0022-5142. (John Wiley & Sons Ltd.)Anthocyanin exts. are increasingly used as food colorants. So far, anthocyanins have not been broadly used in foods and beverages, since they are not as stable as synthetic dyes. Copigmentation between anthocyanins and copigments is the main color-stabilizing mechanism. The process of copigmentation between isolated acylated anthocyanin and rutin, QSA or baicalin has been obsd. using UV-vis spectrophotometry. The thermodn. parameters were correlated to the structure and position of the substituents in the interacting mols. The acylated anthocyanin was isolated from cultivars of Scutellaria baicalensis Georgi flowers and purified by column chromatog. by our own method and has been identified by 1H-/13C-NMR spectroscopy and electrospray mass spectrometry as delphinidin-3-O-(6-O-malonyl)-β-D-glucopyranosyl-5-O-β-D-glucopyranoside.
- 105Ferreira da Silva, P.; Lima, J. C.; Freitas, A. A.; Shimizu, K.; Maçanita, A. L.; Quina, F. H. Charge-Transfer Complexation as a General Phenomenon in the Copigmentation of Anthocyanins J. Phys. Chem. A 2005, 109 (32) 7329– 7338 DOI: 10.1021/jp052106sGoogle Scholar105Charge-Transfer Complexation as a General Phenomenon in the Copigmentation of AnthocyaninsFerreira da Silva, Palmira; Lima, Joao C.; Freitas, Adilson A.; Shimizu, Karina; Macanita, Antonio L.; Quina, Frank H.Journal of Physical Chemistry A (2005), 109 (32), 7329-7338CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Color intensification of anthocyanin solns. in the presence of natural polyphenols (copigmentation) is re-interpreted in terms of charge transfer from the copigment to the anthocyanin. Flavylium cations are shown to be excellent electron acceptors (Ered ≈ -0.3 V vs SCE). It is also demonstrated, for a large series of anthocyanin-copigment pairs, that the std. Gibbs free energy of complex formation decreases linearly with EAAnthoc-IPCop, the difference between the electron affinity of the anthocyanin, EAAnthoc, and the ionization potential of the copigment, IPCop. Based on this correlation, copigmentation strengths of potential candidates for copigments can be predicted.
- 106Galland, S.; Mora, N.; Abert-Vian, M.; Rakotomanomana, N.; Dangles, O. Chemical Synthesis of Hydroxycinnamic Acid Glucosides and Evaluation of Their Ability To Stabilize Natural Colors via Anthocyanin Copigmentation J. Agric. Food Chem. 2007, 55 (18) 7573– 7579 DOI: 10.1021/jf071205vGoogle Scholar106Chemical synthesis of hydroxycinnamic acid glucosides and evaluation of their ability to stabilize natural colors via anthocyanin copigmentationGalland, Stephanie; Mora, Nathalie; Abert-Vian, Maryline; Rakotomanomana, Njara; Dangles, OlivierJournal of Agricultural and Food Chemistry (2007), 55 (18), 7573-7579CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)This work describes the chem. synthesis of O-aryl-β-D-glucosides and 1-O-β-D-glucosyl esters of hydroxycinnamic acids. In particular, O-aryl-β-D-glucosides were efficiently prepd. via a simple diastereoselective glycosylation procedure using phase transfer conditions. Despite the lability of its ester linkage, 1-O-β-D-caffeoylglucose could also be obtained using a Lewis acid catalyzed glycosylation step and a set of protective groups that can be removed under neutral conditions. Hydroxycinnamic acid O-aryl-β-D-glucosides were then quant. investigated for their affinity for the naturally occurring anthocyanin malvin (pigment). Formation of the π-stacking mol. complexes (copigmentation) was characterized in terms of binding consts. and enthalpy and entropy changes. The glucosyl moiety did not significantly alter these thermodn. parameters, in line with a binding process solely involving the polyphenolic nuclei.
- 107Alluis, B.; Dangles, O. Quercetin (=2-(3,4-Dihydroxyphenyl)-3,5,7-Trihydroxy-4H-1-Benzopyran-4-One) Glycosides and Sulfates: Chemical Synthesis, Complexation, and Antioxidant Properties Helv. Chim. Acta 2001, 84 (5) 1133– 1156 DOI: 10.1002/1522-2675(20010516)84:5<1133::AID-HLCA1133>3.3.CO;2-QGoogle ScholarThere is no corresponding record for this reference.
- 108Nave, F.; Brás, N. F.; Cruz, L.; Teixeira, N.; Mateus, N.; Ramos, M. J.; Di Meo, F.; Trouillas, P.; Dangles, O.; De Freitas, V. Influence of a Flavan-3-Ol Substituent on the Affinity of Anthocyanins (Pigments) toward Vinylcatechin Dimers and Proanthocyanidins (Copigments) J. Phys. Chem. B 2012, 116 (48) 14089– 14099 DOI: 10.1021/jp307782yGoogle Scholar108Influence of a Flavan-3-ol Substituent on the Affinity of Anthocyanins (Pigments) toward Vinylcatechin Dimers and Proanthocyanidins (Copigments)Nave, Frederico; Bras, Natercia F.; Cruz, Luis; Teixeira, Natercia; Mateus, Nuno; Ramos, Maria J.; Di Meo, Florent; Trouillas, Patrick; Dangles, Olivier; De Freitas, VictorJournal of Physical Chemistry B (2012), 116 (48), 14089-14099CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The aim of this study is to investigate interactions possibly taking place in red wine between three flavanols (copigments, CP), i.e., two epimeric vinylcatechin dimers (CP1 and CP2) and catechin dimer B3 (CP3), and a specific pigment resulting from the condensation between the main grape anthocyanin malvidin 3-O-glucoside (oenin) and catechin, catechin-(4→8)-oenin. By comparison with our previous work on oenin itself, the influence of the catechin moiety of the anthocyanin in the binding is established. The thermodn. parameters show that both vinylcatechin dimers exhibit a higher affinity for catechin-(4→8)-oenin, in comparison with proanthocyanidin B3, as previously obsd. with oenin. However, the corresponding binding consts. are weaker, probably due to steric hindrance in the anthocyanin brought by the flavanol nucleus. Consequently, catechin-(4→8)-oenin should be much less stabilized by copigmentation in hydroalcoholic soln. than oenin. Quantum mechanics and mol. dynamics simulations are also performed to interpret the binding data, to specify the relative arrangement of the pigment and copigment mols. within the complexes, and to interpret their absorption properties in the visible range.
- 109Dangles, O.; Elhajji, H. Synthesis of 3-Methoxy-and 3-B-D-Glucopyranosy1oxy) Flavylium Ions. Influence of the Flavylium Substitution Pattern on the Reactivity of Anthocyanins in Aqueous Solution Helv. Chim. Acta 1994, 77 (6) 1595– 1610 DOI: 10.1002/hlca.19940770616Google ScholarThere is no corresponding record for this reference.
- 110Cruz, L.; Brás, N. F.; Teixeira, N.; Mateus, N.; Ramos, M. J.; Dangles, O.; De Freitas, V. Vinylcatechin Dimers Are Much Better Copigments for Anthocyanins than Catechin Dimer Procyanidin B3 J. Agric. Food Chem. 2010, 58 (5) 3159– 3166 DOI: 10.1021/jf9037419Google Scholar110Vinylcatechin Dimers Are Much Better Copigments for Anthocyanins than Catechin Dimer Procyanidin B3Cruz, Luis; Bras, Natercia F.; Teixeira, Natercia; Mateus, Nuno; Joao Ramos, Maria; Dangles, Olivier; De Freitas, VictorJournal of Agricultural and Food Chemistry (2010), 58 (5), 3159-3166CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The binding consts. (K) for the interaction of three copigments (CP), two epimeric vinylcatechin dimers (CP1 and CP2), and catechin dimer B3 (CP3) with two pigments, malvidin-3-glucoside (oenin) and malvidin-3,5-diglucoside (malvin), were detd. The K values clearly show that both vinylcatechin dimers have much higher affinity for oenin and malvin than dimer B3: KCP2 > KCP1 » KCP3. Quantum mechanics and mol. dynamics calcns. were also performed to interpret the binding data and specify the relative arrangement of the pigment and copigment mols. within the complexes.
- 111Malien-Aubert, C.; Dangles, O.; Amiot, M. J. Influence of Procyanidins on the Color Stability of Oenin Solutions J. Agric. Food Chem. 2002, 50 (11) 3299– 3305 DOI: 10.1021/jf011392bGoogle Scholar111Influence of Procyanidins on the Color Stability of Oenin SolutionsMalien-Aubert, Celine; Dangles, Olivier; Amiot, Marie JosepheJournal of Agricultural and Food Chemistry (2002), 50 (11), 3299-3305CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The aim of the present work was to specify the influence of the polymn. degree on the color stability of anthocyanins using model solns. under higher thermal conditions simulating rapid food aging. Results showed that an increase in polymeric degree improves the color stability of oenin. Solns. contg. a catechin tetramer, purified from brown rice, displayed a remarkable stability. Flavanols as monomers, (+)-catechin and (-)-epicatechin, appeared to decrease stability with the formation of a xanthylium salt leading to yellowish solns. For the dimers, procyanidin B2 and B3, different behaviors on red color stability have been obsd. corresponding to their different susceptibility to cleavage upon heating. In the presence of the trimeric procyanidin C2, the red color appeared more stable. However, the HPLC chromatograms showed a decrease in the amplitude of the peaks of oenin and procyanidin C2. Concomitantly, a new peak appeared with a maximal absorption in the red region. This newly formed pigment probably came from the condensation of oenin and procyanidin C2.
- 112Benesi, H. A.; Hildebrand, J. H. A Spectrophotometric Investigation of the Interaction of Iodine with Aromatic Hydrocarbons J. Am. Chem. Soc. 1949, 71 (8) 2703– 2707 DOI: 10.1021/ja01176a030Google Scholar112A spectrophotometric investigation of the interaction of iodine with aromatic hydrocarbonsBenesi, H. A.; Hildebrand, J. H.Journal of the American Chemical Society (1949), 71 (), 2703-7CODEN: JACSAT; ISSN:0002-7863.cf. C.A. 42, 8172g. The absorption spectra of I2 in C6H5CF3, benzene, toluene, o- and p-xylene, and mesitylene were measured in the region 270-700 mμ. In the visible region, the absorption peaks of these solns. showed moderate shifts toward shorter wave lengths in the order listed above. With the exception of I2 in C6H5CF3, each of the aromatic hydrocarbon solns. had an intense absorption band in the ultraviolet region which was shown to be characteristic of a complex contg. one I2 and one aromatic hydrocarbon mol. The equil. between I2 and the aromatic hydrocarbons was investigated in the neutral solvents CCl4 and C7H14, and the results show that the iodine-mesitylene complex is more stable than the iodine-benzene complex. These findings are strong evidence for an acid-base interaction between I2 and aromatic hydrocarbons. Absorption measurements also were made of I2 in CCl4, CS2, C7H14, Et2O, Me2CO, and 1,1-dichloroethane. No absorption bands analogous to those of the aromatic hydrocarbon solns. were found in the region 270-400 mμ.
- 113Dangles, O.; Brouillard, R. Polyphenol Interactions. The Copigmentation Case: Thermodynamic Data from Temperature Variation and Relaxation Kinetics. Medium Effect Can. J. Chem. 1992, 70 (8) 2174– 2189 DOI: 10.1139/v92-273Google ScholarThere is no corresponding record for this reference.
- 114Di Meo, F.; Sancho Garcia, J. C.; Dangles, O.; Trouillas, P. Highlights on Anthocyanin Pigmentation and Copigmentation: A Matter of Flavonoid Π-Stacking Complexation To Be Described by DFT-D J. Chem. Theory Comput. 2012, 8 (6) 2034– 2043 DOI: 10.1021/ct300276pGoogle Scholar114Highlights on Anthocyanin Pigmentation and Copigmentation: A Matter of Flavonoid π-Stacking Complexation To Be Described by DFT-DDi Meo, Florent; Sancho Garcia, Juan Carlos; Dangles, Olivier; Trouillas, PatrickJournal of Chemical Theory and Computation (2012), 8 (6), 2034-2043CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Anthocyanidins are a class of π-conjugated systems responsible for red, blue, and purple colors of plants. They exhibit the capacity of aggregation in the presence of other natural compds. including flavonols. Such complexations induce color modulation in plants, which is known as copigmentation. It is largely driven by π-interactions existing between pigments and copigments. In this work, the energies of copigmentation-complexation and self-assocn. are systematically evaluated for an anthocyanidin/flavonol couple prototype (3-O-methylcyanidin/quercetin). To describe noncovalent interactions, DFT-D appears mandatory to reach a large accuracy. Due to the chem. complexity of this phenomenon, we also aim at assessing the relevance of both B3P86-D2 and ωB97X-D functionals. The benchmarking has shown that B3P86-D2 possesses enough accuracy when dealing with π-π interactions with respect to both spin component scaled Moller-Plesset second-order perturbation theory post Hartree-Fock method and exptl. data. UV-vis absorption properties are then evaluated with time-dependent DFT for the different complexes. The use of range-sepd. hybrid functionals, such as ωB97X-D, helped to correctly disentangle and interpret the origin of the UV-vis exptl. shifts attributed to the subtle copigmentation phenomenon.
- 115Awika, J. M. Behavior of 3-Deoxyanthocyanidins in the Presence of Phenolic Copigments Food Res. Int. 2008, 41 (5) 532– 538 DOI: 10.1016/j.foodres.2008.03.002Google Scholar115Behavior of 3-deoxyanthocyanidins in the presence of phenolic copigmentsAwika, Joseph M.Food Research International (2008), 41 (5), 532-538CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)Anthocyanin stability and color intensity are generally improved in the presence of copigments in moderately acidic environments. The 3-deoxyanthocyanins on the other hand are fairly stable to color loss due to change in pH. It is unknown, therefore, how they behave in the presence of copigments. We studied the effects of common phenolic copigments, tannic, ferulic, and O-coumaric acids, and rutin on behavior and stability of six 3-deoxyanthocyandins over 4.5 mo. Tannic and ferulic acid produced the most significant bathochromic shift, whereas rutin had no bathochromic effect. None of the copigments produced a significant hyperchromic shift with the pigments, implying colored species of the pigments were predominant under conditions used. Ferulic and tannic acids were the most effective at improving color stability of 5-hydroxylated pigments, whereas tannic acid and rutin improved the stability of 5-methoxylated pigments the most. Substitution at C-5 was key to overall behavior of the 3-deoxyanthocyanidins.
- 116El Hajji, H.; Dangles, O.; Figueiredo, P.; Brouillard, R. 3′-(β-D-Glycopyranosyloxy) Flavylium Ions: Synthesis and Investigation of Their Properties in Aqueous Solution. Hydrogen Bonding as a Mean of Colour Variation Helv. Chim. Acta 1997, 80 (2) 398– 413 DOI: 10.1002/hlca.19970800206Google ScholarThere is no corresponding record for this reference.
- 117Limón, P. M.; Gavara, R.; Pina, F. Thermodynamics and Kinetics of Cyanidin 3-Glucoside and Caffeine Copigments J. Agric. Food Chem. 2013, 61 (22) 5245– 5251 DOI: 10.1021/jf4006643Google ScholarThere is no corresponding record for this reference.
- 118Melo, M. J.; Moncada, M. C.; Pina, F. On the Red Colour of Raspberry (Rubus Idaeus) Tetrahedron Lett. 2000, 41 (12) 1987– 1991 DOI: 10.1016/S0040-4039(00)00080-0Google ScholarThere is no corresponding record for this reference.
- 119Gierschner, J.; Lüer, L.; Milián-Medina, B.; Oelkrug, D.; Egelhaaf, H.-J. Highly Emissive H-Aggregates or Aggregation-Induced Emission Quenching? The Photophysics of All-Trans Para-Distyrylbenzene J. Phys. Chem. Lett. 2013, 4 (16) 2686– 2697 DOI: 10.1021/jz400985tGoogle Scholar119Highly Emissive H-Aggregates or Aggregation-Induced Emission Quenching? The Photophysics of All-Trans para-DistyrylbenzeneGierschner, Johannes; Luer, Larry; Milian-Medina, Begona; Oelkrug, Dieter; Egelhaaf, Hans-JoachimJournal of Physical Chemistry Letters (2013), 4 (16), 2686-2697CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The present Perspective critically reexamines the photophysics of para-distyrylbenzene (DSB) as a prototype of herringbone-arranged H-aggregates to resolve the apparent contradiction of the frequently reported aggregation-induced emission quenching in H-aggregates on one side and highly emissive DSB crystals on the other and discusses the signatures and fate of excitons in single- and polycryst. samples, including size and polarization effects.
- 120Gierschner, J.; Park, S. Y. Luminescent Distyrylbenzenes: Tailoring Molecular Structure and Crystalline Morphology J. Mater. Chem. C 2013, 1 (37) 5818– 5832 DOI: 10.1039/c3tc31062kGoogle Scholar120Luminescent distyrylbenzenes: tailoring molecular structure and crystalline morphologyGierschner, Johannes; Park, Soo YoungJournal of Materials Chemistry C: Materials for Optical and Electronic Devices (2013), 1 (37), 5818-5832CODEN: JMCCCX; ISSN:2050-7534. (Royal Society of Chemistry)A review. The last few years have seen a steady increase in small mol. based conjugated materials, which promise innovative (opto)electronic applications. This requires however a systematic understanding of structure-property relationships, which can only be achieved via libraries of structurally well-defined single cryst. materials based on systematically designed mol. structures. In this feature article, we are presenting structure-property relationships of functionalized distyrylbenzene (DSB), which is one of the most extensively investigated π-conjugated mol. materials. This will provide a general insight into the specific implications of intermol. arrangements on their solid state optoelectronic properties, discussing H- vs. J-aggregation, herringbone vs. π-stacks, the occurrence of excimers, size effects, and polycrystallinity. The systematic insight into DSB functionalization will then suggest pathways towards targeted mol. design strategies, with special focus on the cyano-vinylene motif (DCS materials) which allows for highly fluorescence solid state samples due to synergetic packing effects promoted by its twist elasticity and secondary bonding interaction. Finally, recent advances in the application of DSB-/DCS-based materials are shortly reviewed.
- 121Gonnet, J.-F. Colour Effects of Co-Pigmentation of Anthocyanins revisited—1. A Colorimetric Definition Using the CIELAB Scale Food Chem. 1998, 63 (3) 409– 415 DOI: 10.1016/S0308-8146(98)00053-3Google ScholarThere is no corresponding record for this reference.
- 122Gordillo, B.; Rodríguez-Pulido, F. J.; Escudero-Gilete, M. L.; González-Miret, M. L.; Heredia, F. J. Comprehensive Colorimetric Study of Anthocyanic Copigmentation in Model Solutions. Effects of pH and Molar Ratio J. Agric. Food Chem. 2012, 60 (11) 2896– 2905 DOI: 10.1021/jf2046202Google Scholar122Comprehensive Colorimetric Study of Anthocyanic Copigmentation in Model Solutions. Effects of pH and Molar RatioGordillo, Belen; Rodriguez-Pulido, Francisco J.; Escudero-Gilete, M. Luisa; Gonzalez-Miret, M. Lourdes; Heredia, Francisco J.Journal of Agricultural and Food Chemistry (2012), 60 (11), 2896-2905CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)New colorimetric variables were defined in the uniform CIELAB color space to assess the quant. and qual. color changes induced by copigmentation and their incidence on visual perception. The copigmentation process was assayed in model solns. between malvidin 3-glucoside and 3 phenolic compds. (catechin, epicatechin, and caffeic acid) as a function of the pH and the pigment/copigment molar ratio. Along the pH variation, the greatest magnitude of copigmentation was obtained at pH 3.0, being significantly higher with epicatechin and caffeic acid. At high acidic pH, the main contribution of copigmentation to the total color was qual., whereas between pH 2.0 and 4.0, the main colorimetric contribution was quant. The contribution of epicatechin and caffeic acid to the color changes was more marked for the quant. characteristics. On contrast, particularly at higher pH values, the qual. contribution was more important in catechin copigmented solns. Increasing copigment concn. induced perceptible color changes at molar ratios higher than 1:2, consisting in a bluish and darkening effect of the anthocyanin solns. Among the different CIELAB attributes, hue difference was the best correlated parameter with the increase of copigment concn., proving the relevance of this physicochem. phenomenon on the qual. changes of anthocyanin color.
- 123Gonnet, J.-F. Colour Effects of Co-Pigmentation of Anthocyanins revisited—2.A Colorimetric Look at the Solutions of Cyanin Co-Pigmented Byrutin Using the CIELAB Scale Food Chem. 1999, 66 (3) 387– 394 DOI: 10.1016/S0308-8146(99)00088-6Google ScholarThere is no corresponding record for this reference.
- 124Xu, H.; Liu, X.; Yan, Q.; Yuan, F.; Gao, Y. A Novel Copigment of Quercetagetin for Stabilization of Grape Skin Anthocyanins Food Chem. 2015, 166, 50– 55 DOI: 10.1016/j.foodchem.2014.05.125Google Scholar124A novel copigment of quercetagetin for stabilization of grape skin anthocyaninsXu, Honggao; Liu, Xuan; Yan, Qiuli; Yuan, Fang; Gao, YanxiangFood Chemistry (2015), 166 (), 50-55CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)The thermal and light stability of grape skin anthocyanins combined with quercetagetin was investigated at designed pH values of 3, 4 and 5. The molar ratios of anthocyanins to quercetagetin were 1:10, 1:20 and 1:40 for thermally treatment at 70 °C, 80 °C and 90 °C, resp., and the ratios were tested at 5:1, 1:1, 1:5 and 1:10 in the light exposure expts. The degrdn. reaction of anthocyanins in the presence of quercetagetin followed the first-order kinetic model. The half-life (t1/2) of anthocyanins was extended significantly with the increase of quercetagetin concn. (p < 0.05). The total color difference values (ΔE*) for the anthocyanin solns. with quercetagetin were smaller than those without copigment under the same exptl. conditions (pH and light exposure time). Compared with epigallocatechin gallate (EGCG), tea polyphenols (TP), myricitrin and rutin, quercetagetin was the most effective copigment to stabilize grape skin anthocyanins.
- 125Shao, P.; Zhang, J.; Fang, Z.; Sun, P. Complexing of Chlorogenic Acid with B-Cyclodextrins: Inclusion Effects, Antioxidative Properties and Potential Application in Grape Juice Food Hydrocolloids 2014, 41, 132– 139 DOI: 10.1016/j.foodhyd.2014.04.003Google ScholarThere is no corresponding record for this reference.
- 126Mistry, T. V.; Cai, Y.; Lilley, T. H.; Haslam, E. Polyphenol Interactions. Part 5. Anthocyanin Co-Pigmentation J. Chem. Soc., Perkin Trans. 2 1991, 8) 1287– 1296 DOI: 10.1039/p29910001287Google Scholar126Polyphenol interactions. Part 5. Anthocyanin copigmentationMistry, Tarankumar V.; Cai, Ya; Lilley, Terence H.; Haslam, EdwinJournal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1991), (8), 1287-96CODEN: JCPKBH; ISSN:0300-9580.In aq. media, anthocyanins undergo several structural transformations and exist in a series of equil. between carbinol-base, flavylium cation, quinonoidal anhydro-base and chalcone forms. A detailed interpretation of the 1H-NMR (400 MHz) spectra of malvin in D2O is given for the first time in the context of these equil. The phenomenon of copigmentation is reviewed and the efficacy of various phenolic flavonoids, galloyl and hexahydroxycinnamyl esters as natural copigments is detd. quant. DNA, RNA and ATP also act as effective copigments for malvin in the form the flavylium ion but both caffeine and theophylline preferentially stabilize the quinonoidal-base forms of the anthocyanin. At pH 3.5 to 7.0, they give rise to stable violet through to blue and finally green colors. 1H-NMR studies of all these intermol. copigmentation reactions are reported and the phenomenon is provisionally interpreted in terms of hydrophobically reinforced π-π stacking of anthocyanin and copigment mols. in the aq. environment.
- 127Son, T.-D.; Chachaty, C. Nucleoside Conformations: XIV. Conformation of Adenosine Monophosphates in Aqueous Solution by Proton Magnetic Resonance Spectroscopy Biochim. Biophys. Acta, Nucleic Acids Protein Synth. 1974, 335 (1) 1– 13 DOI: 10.1016/0005-2787(74)90234-2Google ScholarThere is no corresponding record for this reference.
- 128Leydet, Y.; Gavara, R.; Petrov, V.; Diniz, A. M.; Jorge Parola, A.; Lima, J. C.; Pina, F. The Effect of Self-Aggregation on the Determination of the Kinetic and Thermodynamic Constants of the Network of Chemical Reactions in 3-Glucoside Anthocyanins Phytochemistry 2012, 83, 125– 135 DOI: 10.1016/j.phytochem.2012.06.022Google Scholar128The effect of self-aggregation on the determination of the kinetic and thermodynamic constants of the network of chemical reactions in 3-glucoside anthocyaninsLeydet, Yoann; Gavara, Raquel; Petrov, Vesselin; Diniz, Ana M.; Parola, A. Jorge; Lima, Joao C.; Pina, FernandoPhytochemistry (Elsevier) (2012), 83 (), 125-135CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Ltd.)The six most common 3-glucoside anthocyanins, pelargonidin-3-glucoside, peonidin-3-glucoside, delphinidin-3-glucoside, malvidin-3-glucoside, cyanidin-3-glucoside and petunidin-3-glucoside were studied in great detail by NMR, UV-vis absorption and stopped flow. For each anthocyanin, the thermodn. and kinetic consts. of the network of chem. reactions were calcd. at different anthocyanin concn., from 6 × 10-6 M up to 8 × 10-4 M; an increasing of the flavylium cation acidity const. to give quinoidal base and a decreasing of the flavylium cation hydration const. to give hemiketal were obsd. by increasing the anthocyanin concn. These effects are attributed to the self-aggregation of the flavylium cation and quinoidal base, which is stronger in the last case. The UV-vis and 1H NMR spectral variations resulting from the increasing of the anthocyanin concn. were discussed in terms of two aggregation models; monomer-dimer and isodesmic, the last one considering the formation of higher order aggregates possessing the same aggregation const. of the dimer. The self-aggregation const. of flavylium cation at pH = 1.0, calcd. by both models increases by increasing the no. of methoxy (-OCH3) or hydroxy (-OH) substituents following the order: myrtillin (2 -OH), oenin (2 -OCH3), 3-OGl-petunidin (1 -OH, 1 -OCH3), kuromanin (1 -OH), 3-OGl-peonidin (1 -OCH3) and callistephin (none). Evidence for flavylium aggregates possessing a shape between J and H was achieved, as well as for the formation of higher order aggregates.
- 129Dimicoli, J. L.; Hélène, C. Complex Formation between Purine and Indole Derivatives in Aqueous Solutions. Proton Magnetic Resonance Studies J. Am. Chem. Soc. 1973, 95 (4) 1036– 1044 DOI: 10.1021/ja00785a008Google ScholarThere is no corresponding record for this reference.
- 130Alluis, B.; Pérol, N.; El Hajji, H.; Dangles, O. Water-Soluble Flavonol (=3Hydroxy2-Phenyl-4H-1-Benzopyran-4-One) Derivatives: Chemical Synthesis, Colouring, and Antioxidant Properties Helv. Chim. Acta 2000, 83 (2) 428– 443 DOI: 10.1002/(SICI)1522-2675(20000216)83:2<428::AID-HLCA428>3.3.CO;2-AGoogle ScholarThere is no corresponding record for this reference.
- 131Hondo, T.; Yoshida, K.; Nakagawa, A.; Kawai, T.; Tamura, H.; Goto, T. Structural Basis of Blue-Colour Development in Flower Petals from Commelina Communis Nature 1992, 358 (6386) 515– 518 DOI: 10.1038/358515a0Google ScholarThere is no corresponding record for this reference.
- 132Shiono, M.; Matsugaki, N.; Takeda, K. Phytochemistry: Structure of the Blue Cornflower Pigment Nature 2005, 436 (7052) 791– 791 DOI: 10.1038/436791aGoogle ScholarThere is no corresponding record for this reference.
- 133Dufour, C.; Sauvaitre, I. Interactions between Anthocyanins and Aroma Substances in a Model System. Effect on the Flavor of Grape-Derived Beverages J. Agric. Food Chem. 2000, 48 (5) 1784– 1788 DOI: 10.1021/jf990877lGoogle Scholar133Interactions between anthocyanins and aroma substances in a model system. Effect on the flavor of grape-derived beveragesDufour, Claire; Sauvaitre, IsabelleJournal of Agricultural and Food Chemistry (2000), 48 (5), 1784-1788CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Evaluation of the sensory quality of wine or grape-derived beverages led the authors to study the interactions between flavors and anthocyanins, the colored family of polyphenols. The flavylium cation-ligand complexation, resulting in copigmentation (rise in pigment visible absorption with a concomitant bathochromic shift), was investigated using visible absorption spectroscopy. Sole volatile phenols were found to markedly interact with malvidin-3,5-O-diglucoside. With series of guaiacyl-derived aroma substances, acyl-substituted ligands proved to be better copigments than alkyl-substituted ones. Assocn. consts. and 1:1 complex stoichiometry were further detd. for several substrates. Decreasing binding to malvin was obsd. for acetosyringone, syringaldehyde, acetovanillone, vanillin, 3,5-dimethoxyphenol, and 4-ethylguaiacol. Addn. of 10% ethanol lowered by one-third the assocn. consts. for malvin-ligand couples and for malvidin-3-O-glucoside with acetosyringone and syringaldehyde. The main driving force was ascribed to hydrophobicity, although this study evidenced an influence of the ligand substitution pattern on copigmentation.
- 134Tuominen, A.; Sinkkonen, J.; Karonen, M.; Salminen, J.-P. Sylvatiins, Acetylglucosylated Hydrolysable Tannins from the Petals of Geranium Sylvaticum Show Co-Pigment Effect Phytochemistry 2015, 115, 239– 251 DOI: 10.1016/j.phytochem.2015.01.005Google Scholar134Sylvatiins, acetylglucosylated hydrolysable tannins from the petals of Geranium sylvaticum show co-pigment effectTuominen, Anu; Sinkkonen, Jari; Karonen, Maarit; Salminen, Juha-PekkaPhytochemistry (Elsevier) (2015), 115 (), 239-251CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Ltd.)Four hydrolysable tannins, named as sylvatiins A (1), B (2), C (3) and D (4), were isolated from the petals of Geranium sylvaticum. On the basis of spectrometric evidence of NMR anal. (1H NMR, 13C NMR, DQF-COSY, TOCSY, NOESY, HSQC and HMBC), CD and ESI-MS/MS, sylvatiins A, B and C were characterized as galloyl glucoses contg. one or two acetylglucoses attached to the 3-OH of the galloyl group, whereas sylvatiin D was found to have a chebulinic acid core contg. acetylglucose attached in a similar way. The potential of these compds. to act as defensive compds. against herbivores was evaluated using the radial diffusion assay that measures the protein pptn. capacity. In addn., the capacity of sylvatiins to act as co-pigments with anthocyanins of G. sylvaticum petals was measured in vitro at different pH values. Sylvatiins A and D maintained efficiently the purple flower color near the natural pH of petal cells. The amt. of sylvatiins was changed according to the flower color; deep purple petals with higher amt. of anthocyanin contained more sylvatiins A and C than whiter petals. It was concluded that G. sylvaticum petal cells may accumulate sylvatiins for intermol. co-pigmentation purposes.
- 135Dangles, O.; Saito, N.; Brouillard, R. Kinetic and Thermodynamic Control of Flavylium Hydration in the Pelargonidin-Cinnamic Acid Complexation. Origin of the Extraordinary Flower Color Diversity of Pharbitis Nil J. Am. Chem. Soc. 1993, 115 (8) 3125– 3132 DOI: 10.1021/ja00061a011Google ScholarThere is no corresponding record for this reference.
- 136Eiro, M. J.; Heinonen, M. Anthocyanin Color Behavior and Stability during Storage: Effect of Intermolecular Copigmentation J. Agric. Food Chem. 2002, 50 (25) 7461– 7466 DOI: 10.1021/jf0258306Google Scholar136Anthocyanin Color Behavior and Stability during Storage: Effect of Intermolecular CopigmentationEiro, Maarit J.; Heinonen, MarinaJournal of Agricultural and Food Chemistry (2002), 50 (25), 7461-7466CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Intermol. copigmentation reactions are significantly responsible for the manifold color expression of fruits, berries, and their products. These reactions were investigated with 5 anthocyanins and 5 phenolic acids acting as copigments. The stability of the pigment-copigment complexes formed was studied during a storage period of 6 mo. The study was conducted using a UV-visible spectrophotometer to monitor the hyperchromic effect and the bathochromic shift of the complexes. The greatest copigmentation reactions took place in malvidin 3-glucoside solns. The strongest copigments for all anthocyanins were ferulic and rosmarinic acids. The immediate reaction of rosmarinic acid with malvidin 3-glucoside resulted in the biggest bathochromic shift (19 nm) and the strongest hyperchromic effect, increasing the color intensity by 260%. The color induced by rosmarinic acid was not very stable. The color intensity of pelargonidin 3-glucoside increased greatly throughout the storage period with the addn. of ferulic and caffeic acids.
- 137Houbiers, C.; Lima, J. C.; Maçanita, A. L.; Santos, H. Color Stabilization of Malvidin 3-Glucoside: Self-Aggregation of the Flavylium Cation and Copigmentation with the Z-Chalcone Form J. Phys. Chem. B 1998, 102 (18) 3578– 3585 DOI: 10.1021/jp972320jGoogle ScholarThere is no corresponding record for this reference.
- 138Elhabiri, M.; Figueiredo, P.; Toki, K.; Figueiredo, P.; Toki, K.; Saito, N.; Brouillard, R.; Figueiredo, P.; Toki, K. Anthocyanin–aluminium and – gallium Complexes in Aqueoussolution J. Chem. Soc., Perkin Trans. 2 1997, 2) 355– 362 DOI: 10.1039/a603851dGoogle Scholar138Anthocyanin-aluminum and -gallium complexes in aqueous solutionElhabiri, M.; Figueiredo, P.; Toki, K.; Saito, N.; Brouillard, R.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1997), (2), 355-362CODEN: JCPKBH; ISSN:0300-9580. (Royal Society of Chemistry)Complexation of aluminum and gallium ions with synthetic anthocyanin models and natural anthocyanins extd. from the blue flowers of Evolvulus pilosus cv 'Blue Daze' and the violet flowers of Matthiola incana has been thoroughly investigated in aq. soln. From UV-VIS spectroscopic data collected at pH 2-5, the presence of complexes, involving not only the colored forms but also the colorless forms of the pigments is demonstrated. A theor. treatment is developed for the calcn. of the corresponding stability consts. The pigments studied throughout this work can be divided into two series, one sharing a cyanidin chromophore and the other a delphinidin one. Within both series, individual pigments are distinguished according to the degree and type of glycosidation and/or acylation. Intramol. effects such as copigmentation of anthocyanin-aluminum complexes and the effect of the presence of a malonyl group on the formation of those complexes are discussed. These results are important to plant pigmentation and, for instance, a narrow pH domain in which color amplification due to complexation is at a max. has been found.
- 139Al Bittar, S.; Mora, N.; Loonis, M.; Dangles, O. Chemically Synthesized Glycosides of Hydroxylated Flavylium Ions as Suitable Models of Anthocyanins: Binding to Iron Ions and Human Serum Albumin, Antioxidant Activity in Model Gastric Conditions Molecules 2014, 19 (12) 20709– 20730 DOI: 10.3390/molecules191220709Google ScholarThere is no corresponding record for this reference.
- 140Kasha, M.; Rawls, H. R.; Ashraf El-Bayoumi, M. The Exciton Model in Molecular Spectroscopy Pure Appl. Chem. 1965, 11 (3–4) 371– 392 DOI: 10.1351/pac196511030371Google Scholar140Exciton model in molecular spectroscopyKasha, Michael; Rawls, Henry R.; El-Bayoumi, M. AshrafPure and Applied Chemistry (1965), 11 (3-4), 371-92CODEN: PACHAS; ISSN:0033-4545.A summary is presented of the application of the mol. exciton model to dimers, trimers, and double and triple mols. In those cases where no significant exciton effect is observable in the singlet-singlet absorption spectrum for the composite mol., the enhancement of lowest triplet state excitation may still be conspicuous and significant.
- 141Yoshida, K.; Toyama, Y.; Kameda, K.; Kondo, T. Contribution of Each Caffeoyl Residue of the Pigment Molecule of Gentiodelphin to Blue Color Development Phytochemistry 2000, 54 (1) 85– 92 DOI: 10.1016/S0031-9422(00)00049-2Google ScholarThere is no corresponding record for this reference.
- 142Yoshida, K.; Kondo, T.; Goto, T. Intramolecular Stacking Conformation of Gentiodelphin, a Diacylated Anthocyanin from Gentiana Makinoi Tetrahedron 1992, 48 (21) 4313– 4326 DOI: 10.1016/S0040-4020(01)80442-7Google ScholarThere is no corresponding record for this reference.
- 143Nerdal, W.; Andersen, Ø. M. Evidence for Self-Association of the Anthocyanin Petanin in Acidified, Methanolic Solution Using Two-Dimensional Nuclear Overhauser Enhancement NMR Experiments and Distance Geometry Calculations Phytochem. Anal. 1991, 2 (6) 263– 270 DOI: 10.1002/pca.2800020606Google Scholar143Evidence for self-association of the anthocyanin petanin in acidified, methanolic solution using two-dimensional nuclear Overhauser enhancement NMR experiments and distance geometry calculationsNerdal, Willy; Andersen, Oeyvind M.Phytochemical Analysis (1991), 2 (6), 263-70CODEN: PHANEL; ISSN:0958-0344.A new mechanism is described for self-assocn. of anthocyanins in soln. Petanin [petunidin 3-O-[6-O-(4-O-E-p-coumaroyl-α-L-rhamnopyranosyl)-β-D-glucopyranoside]-5-O-β-D-glucopuranoside] was studied in acidified methanolic soln. using the nuclear Overhauser enhancement (NOESY) NMR technique. Intra- and inter-mol. NOESY cross-peaks were obsd., and the corresponding proton-proton distance bounds were used in distance geometry calcns. to det. the stacking of the petanin aglycons. The orientation of two self-assocd. petanin aglycons was found to be head-to-tail along both the long and the short aglycon axis. Lack of obsd. NOESY cross-peaks between protons of the coumaroyl group and the aglycon indicated absence of intramol. stacking of the stable petanin mols. Non-coplanarity between the planes of the benzopyrylium and the Ph rings was also indicated.
- 144Nerdal, W.; Andersen, Ø. M. Intermolecular Aromatic Acid Association of an Anthocyanin (petanin) Evidenced by Two-Dimensional Nuclear Overhauser Enhancement Nuclear Magnetic Resonance Experiments and Distance Geometry Calculations Phytochem. Anal. 1992, 3 (4) 182– 189 DOI: 10.1002/pca.2800030408Google Scholar144Intermolecular aromatic acid association of an anthocyanin (petanin) evidence by two-dimensional nuclear Overhauser enhancement nuclear magnetic resonance experiments and distance geometry calculationsNerdal, Willy; Andersen, Oeyvind M.Phytochemical Analysis (1992), 3 (4), 182-9CODEN: PHANEL; ISSN:0958-0344.A new mechanism for the mol. assocn. of the arom acid of an anthocyanin using two-dimensional 1H NOESY expts., is described.
- 145Yoshida, K.; Kondo, T.; Goto, T. Unusually Stable Monoacylated Anthocyanin from Purple Yam Dioscorea Alata Tetrahedron Lett. 1991, 32 (40) 5579– 5580 DOI: 10.1016/0040-4039(91)80088-NGoogle ScholarThere is no corresponding record for this reference.
- 146Mori, M.; Miki, N.; Ito, D.; Kondo, T.; Yoshida, K. Structure of Tecophilin, a Tri-Caffeoylanthocyanin from the Blue Petals of Tecophilaea Cyanocrocus, and the Mechanism of Blue Color Development Tetrahedron 2014, 70 (45) 8657– 8664 DOI: 10.1016/j.tet.2014.09.046Google Scholar146Structure of tecophilin, a tri-caffeoylanthocyanin from the blue petals of Tecophilaea cyanocrocus, and the mechanism of blue color developmentMori, Mihoko; Miki, Naoko; Ito, Daisuke; Kondo, Tadao; Yoshida, KumiTetrahedron (2014), 70 (45), 8657-8664CODEN: TETRAB; ISSN:0040-4020. (Elsevier Ltd.)The pigment, tecophilin, in blue flowers of Tecophilaea cyanocrocus was isolated and the structure was detd. to be 3-O-(6-O-α-L-rhamnopyranosyl-β-D-glucopyranosyl)-7-O-(6-O-(4-O-(2-O-(4-O-β-D-glucopyranosyl-(E)-caffeoyl)-6-O-(4-O-β-D-glucopyranosyl-(E)-caffeoyl)-β-D-glucopyranosyl)-(E)-caffeoyl)-β-D-glucopyranosyl)delphinidin. The reprodn. expt. of the same color as petals according to the results of chem. anal. and measurement of vacuolar pH of blue cells clarified that the blue color solely develops by tecophilin without interaction of metal ions nor co-pigments. 1H NMR anal. and CD spectrum indicate the co-existence of clockwise intermol. self-assocn. of the delphinidin nuclei and intramol. π-π stacking between the chromophore and caffeoyl residues to derive bathochromic shift of the absorption spectrum and stabilize the color by preventing hydration reaction.
- 147Terahara, N.; Callebaut, A.; Ohba, R.; Nagata, T.; Ohnishi-Kameyama, M.; Suzuki, M. Triacylated Anthocyanins from Ajuga Reptans Flowers and Cell Cultures Phytochemistry 1996, 42 (1) 199– 203 DOI: 10.1016/0031-9422(95)00838-1Google ScholarThere is no corresponding record for this reference.
- 148Terahara, N.; Toki, K.; Saito, N.; Honda, T.; Matsui, T.; Osajima, Y. Eight New Anthocyanins, Ternatins C1-C5 and D3 and Preternatins A3 and C4 from Young Clitoria Ternatea Flowers J. Nat. Prod. 1998, 61 (11) 1361– 1367 DOI: 10.1021/np980160cGoogle ScholarThere is no corresponding record for this reference.
- 149Escribano-Bailón, T.; Dangles, O.; Brouillard, R. Coupling Reactions between Flavylium Ions and Catechin Phytochemistry 1996, 41 (6) 1583– 1592 DOI: 10.1016/0031-9422(95)00811-XGoogle Scholar149Coupling reactions between flavylium ions and catechinEscribano-Bailon, Teresa; Dangles, Olivier; Brouillard, RaymondPhytochemistry (1996), 41 (6), 1583-92CODEN: PYTCAS; ISSN:0031-9422. (Elsevier)In order to model natural polymeric pigments present in old red wines, new covalent adducts have been synthesized upon condensation of synthetic flavylium ions (models of anthocyanins) with catechin (model of tannins) in the presence and in the absence of acetaldehyde. These new pigments have been investigated by 1D and 2D NMR, HPLC, FAB-mass and UV-visible spectroscopies and mol. modeling. The two flavylium salts used in this work (3,4'-dimethoxy-7-hydroxyflavylium chloride and 5,7-dihydroxy-3,4'-dimethoxyflavylium chloride) display quite different reactivities toward catechin. The electronic donating effect of the catechin moiety and the formation of noncovalent dimers in acidic aq. or methanolic soln. should be mainly responsible for the improved stability of the flavylium chromophore in the new pigments.
- 150Dueñas, M.; Fulcrand, H.; Cheynier, V. Formation of Anthocyanin–flavanol Adducts in Model Solutions Anal. Chim. Acta 2006, 563 (1–2) 15– 25 DOI: 10.1016/j.aca.2005.10.062Google ScholarThere is no corresponding record for this reference.
- 151Dueñas, M.; Salas, E.; Cheynier, V.; Dangles, O.; Fulcrand, H. UV–Visible Spectroscopic Investigation of the 8,8-Methylmethine Catechin-Malvidin 3-Glucoside Pigments in Aqueous Solution: Structural Transformations and Molecular Complexation with Chlorogenic Acid J. Agric. Food Chem. 2006, 54 (1) 189– 196 DOI: 10.1021/jf0516989Google ScholarThere is no corresponding record for this reference.
- 152Chassaing, S.; Lefeuvre, D.; Jacquet, R.; Jourdes, M.; Ducasse, L.; Galland, S.; Grelard, A.; Saucier, C.; Teissedre, P.-L.; Dangles, O. Physicochemical Studies of New Anthocyano-Ellagitannin Hybrid Pigments: About the Origin of the Influence of Oak C-Glycosidic Ellagitannins on Wine Color Eur. J. Org. Chem. 2010, 2010 (1) 55– 63 DOI: 10.1002/ejoc.200901133Google ScholarThere is no corresponding record for this reference.
- 153Bloor, S. J. Overview of Methods for Analysis and Identification of Flavonoids Methods Enzymol. 2001, 335, 3– 14 DOI: 10.1016/S0076-6879(01)35227-8Google ScholarThere is no corresponding record for this reference.
- 154Dangles, O.; Saito, N.; Brouillard, R. Anthocyanin Intramolecular Copigment Effect Phytochemistry 1993, 34 (1) 119– 124 DOI: 10.1016/S0031-9422(00)90792-1Google ScholarThere is no corresponding record for this reference.
- 155Figueiredo, P.; George, F.; Tatsuzawa, F.; Toki, K.; Saito, N.; Brouillard, R. New Features of Intramolecular Copigmentation by Acylated Anthocyanins Phytochemistry 1999, 51, 125– 132 DOI: 10.1016/S0031-9422(98)00685-2Google Scholar155New features of intramolecular copigmentation by acylated anthocyaninsFigueiredo, Paulo; George, Florian; Tatsuzawa, Fumi; Toki, Kenjiro; Saito, Norio; Brouillard, RaymondPhytochemistry (1999), 51 (1), 125-132CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Science Ltd.)Three series of structurally related anthocyanins, extd. from the red-purple flowers of Dendrobium "Pramot", xLaeliocattleya cv. Mini Purple, Bletilla striata and Phalaenopsis all belonging to the Orchidaceae family and another series extd. from the pink flowers of Senecio cruentus (Compositae) allowed the confirmation of the existence of strong intramol. copigmentation effects. These interactions confer stability to the colored forms of the mols., in a wide range of slightly acidic to neutral aq. media. Moreover, the existence of structural relationships among the four series stressed the different influences exerted by the diverse substituent groups. The existence of a malonylglucoside attached to position 3 of all but three of the mols. put forward a new role for the malonyl residue, in this particular position.
- 156Fleschhut, J.; Kratzer, F.; Rechkemmer, G.; Kulling, S. E. Stability and Biotransformation of Various Dietary Anthocyanins in Vitro Eur. J. Nutr. 2006, 45 (1) 7– 18 DOI: 10.1007/s00394-005-0557-8Google Scholar156Stability and biotransformation of various dietary anthocyanins in vitroFleschhut, Jens; Kratzer, Frank; Rechkemmer, Gerhard; Kulling, Sabine E.European Journal of Nutrition (2006), 45 (1), 7-18CODEN: EJNUFZ; ISSN:1436-6207. (Steinkopff Verlag)Background Anthocyanins, which are found in high concns. in fruit and vegetable, may play a beneficial role in retarding or reversing the course of chronic degenerative diseases. However, little is known about the biotransformation and the metab. of anthocyanins so far. Aim of the study The aim of the study was to investigate possible transformation pathways of anthocyanins by human fecal microflora and by rat liver microsomes as a source of cytochrome P 450 enzymes as well as of glucuronyltransferases. Methods Pure anthocyanins, an aq. ext. of red radish as well as the assumed degrdn. products were incubated with human fecal suspension. The incubation mixts. were purified by solid-phase extn. and analyzed by HPLC/DAD/MS and GC/MS. Quantification was done by the external std. method. Furthermore the biotransformation of anthocyanins by incubation with rat liver microsomes in the presence of the cofactor NADPH (as a model for the phase I oxidn.) and in the presence of activated glucuronic acid (as a model for the phase II glucuronidation) was investigated. Results Glycosylated and acylated anthocyanins were rapidly degraded by the intestinal microflora after anaerobic incubation with a human fecal suspension. The major stable products of anthocyanin degrdn. are the corresponding phenolic acids derived from the B-ring of the anthocyanin skeleton. Anthocyanins were not metabolized by cytochrome P 450 enzymes, neither hydroxylated nor demethylated. However they were glucuronidated by rat liver microsomes to several products. Conclusions The gut microflora seem to play an important role in the biotransformation of anthocyanins. A rapid degrdn. could be one major reason for the poor bioavailability of anthocyanins in pharmacokinetic studies described so far in the literature. The formation of phenolic acids as the major stable degrdn. products gives an important hint to the fate of anthocyanins in vivo.
- 157Lopes, P.; Richard, T.; Saucier, C.; Teissedre, P.-L.; Monti, J.-P.; Glories, Y. Anthocyanone A: A Quinone Methide Derivative Resulting from Malvidin 3- O -Glucoside Degradation J. Agric. Food Chem. 2007, 55 (7) 2698– 2704 DOI: 10.1021/jf062875oGoogle ScholarThere is no corresponding record for this reference.
- 158Sadilova, E.; Carle, R.; Stintzing, F. C. Thermal Degradation of Anthocyanins and Its Impact on Color Andin Vitro Antioxidant Capacity Mol. Nutr. Food Res. 2007, 51 (12) 1461– 1471 DOI: 10.1002/mnfr.200700179Google Scholar158Thermal degradation of anthocyanins and its impact on color and in vitro antioxidant capacitySadilova, Eva; Carle, Reinhold; Stintzing, Florian C.Molecular Nutrition & Food Research (2007), 51 (12), 1461-1471CODEN: MNFRCV; ISSN:1613-4125. (Wiley-VCH Verlag GmbH & Co. KGaA)The aim of the current study was to thoroughly investigate the structural changes of anthocyanins at pH 3.5 in purified fractions from black carrot, elderberry and strawberry heated over 6 h at 95°C. Degrdn. products were monitored by HPLC-DAD-MS3 to elucidate the prevailing degrdn. pathways. In addn., alterations of color and antioxidant properties obsd. upon heating were scrutinized. Most interestingly, the degrdn. pathways at pH 3.5 were found to differ from those at pH 1. Among others, chalcone glycosides were detected at 320 nm in heat-treated elderberry and strawberry pigment isolates, and opening of the pyrylium ring initiated anthocyanin degrdn. In the case of acylated anthocyanins, acyl-glycoside moieties were split off from the flavylium backbone, first. Finally, for all pigment isolates, phenolic acids and phloroglucinaldehyde were the terminal degrdn. products as remainders of the B- and A-ring, resp. Maximum and min. antioxidant stabilizing capacities were found in black carrot and strawberry, resp., which was explained by the high degree of acylation in the former. After heating, decline of trolox equiv. antioxidant capacity (TEAC) was obsd. in all samples, which was attributed to both anthocyanins and their colorless degrdn. products following thermal exposure. As deduced from the ratio of TEAC value and anthocyanin content, the loss of anthocyanin bioactivity could not be compensated by the antioxidant capacity of newly formed colorless phenolics upon heating.
- 159Yang, J.; Hu, W.; Usvyat, D.; Matthews, D.; Schütz, M.; Chan, G. K.-L. Ab Initio Determination of the Crystalline Benzene Lattice Energy to Sub-Kilojoule/mol Accuracy Science 2014, 345 (6197) 640– 643 DOI: 10.1126/science.1254419Google Scholar159Ab initio determination of the crystalline benzene lattice energy to sub-kilojoule/mole accuracyYang, Jun; Hu, Weifeng; Usvyat, Denis; Matthews, Devin; Schuetz, Martin; Chan, Garnet Kin-LicScience (Washington, DC, United States) (2014), 345 (6197), 640-643CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Computation of lattice energies to an accuracy sufficient to distinguish polymorphs is a fundamental bottleneck in crystal structure prediction. For the lattice energy of the prototypical benzene crystal, we combined the quantum chem. advances of the last decade to attain sub-kilojoule per mol accuracy, an order-of-magnitude improvement in certainty over prior calcns. that necessitates revision of the exptl. extrapolation to 0 K. Our computations reveal the nature of binding by improving on previously inaccessible or inaccurate multibody and many-electron contributions and provide revised ests. of the effects of temp., vibrations, and relaxation. Our demonstration raises prospects for definitive first-principles resoln. of competing polymorphs in mol. crystal structure prediction.
- 160Cavalcanti, R. N.; Santos, D. T.; Meireles, M. A. A. Non-Thermal Stabilization Mechanisms of Anthocyanins in Model and Food systems—An Overview Food Res. Int. 2011, 44 (2) 499– 509 DOI: 10.1016/j.foodres.2010.12.007Google Scholar160Non-thermal stabilization mechanisms of anthocyanins in model and food systems-An overviewCavalcanti, Rodrigo N.; Santos, Diego T.; Meireles, Maria Angela A.Food Research International (2011), 44 (2), 499-509CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)A review. Phenolic compds. are part of the secondary metab. of plants and are of great importance for their survival in unfavorable environment. A class of phenolic compds. easily found in the Plant Kingdom, is anthocyanins, a flavonoid category. They are water-sol. pigments that confer the bright red, blue, and purple colors of fruits and vegetables and promote several health benefits due to their diverse biol. activities. Different factors affect the color and stability of these compds. including pH, temp., light, presence of copigments, self-assocn., metallic ions, enzymes, oxygen, ascorbic acid, sugar, among others. For this reason many studies have been conducted with the aim to increase the stability of these substances. Therefore, the present review highlights studies on the stabilization of anthocyanins and presents latent anthocyanin stabilization mechanisms and demonstrates the potentiality of the main techniques used: assocn. and encapsulation.
- 161Rodrigues, R. F.; Ferreira da Silva, P.; Shimizu, K.; Freitas, A. A.; Kovalenko, S. A.; Ernsting, N. P.; Quina, F. H.; Maçanita, A. Ultrafast Internal Conversion in a Model Anthocyanin-Polyphenol Complex: Implications for the Biological Role of Anthocyanins in Vegetative Tissues of Plants Chem. - Eur. J. 2009, 15 (6) 1397– 1402 DOI: 10.1002/chem.200801207Google Scholar161Ultrafast internal conversion in a model anthocyanin-polyphenol complex: implications for the biological role of anthocyanins in vegetative tissues of plantsRodrigues, Rita Franca; Ferreira da Silva, Palmira; Shimizu, Karina; Freitas, Adilson A.; Kovalenko, Sergey A.; Ernsting, Nikolaus P.; Quina, Frank H.; Macanita, AntonioChemistry - A European Journal (2009), 15 (6), 1397-1402CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The red flavylium cations of anthocyanins form ground-state charge-transfer complexes with several naturally occurring electron-donor copigments, such as hydroxylated flavones and hydroxycinnamic or benzoic acids. Excitation of the 7-methoxy-4-methyl-flavylium-protocatechuic acid complex results in ultrafast (240 fs) internal conversion to the ground state of the complex by way of a low-lying charge-transfer state. Thus, both uncomplexed anthocyanins, whose excited state decays by fast (5-20 ps) excited-state proton transfer, and anthocyanin-copigment complexes have highly efficient mechanisms of deactivation that are consistent with the proposed protective role of anthocyanins against excess solar radiation in the vegetative tissues of plants.
- 162Ferreira da Silva, P.; Paulo, L.; Barbafina, A.; Elisei, F.; Quina, F. H.; Maçanita, A. L. Photoprotection and the Photophysics of Acylated Anthocyanins Chem. - Eur. J. 2012, 18 (12) 3736– 3744 DOI: 10.1002/chem.201102247Google Scholar162Photoprotection and the Photophysics of Acylated AnthocyaninsFerreira da Silva, Palmira; Paulo, Luisa; Barbafina, Adrianna; Eisei, Fausto; Quina, Frank H.; Macanita, Antonio L.Chemistry - A European Journal (2012), 18 (12), 3736-3744CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The proposed role of anthocyanins in protecting plants against excess solar radiation is consistent with the occurrence of ultrafast (5-25 ps) excited-state proton transfer as the major deexcitation pathway of these mols. However, because natural anthocyanins absorb mainly in the visible region of the spectra, with only a narrow absorption band in the UV-B region, this highly efficient deactivation mechanism would essentially only protect the plant from visible light. On the other hand, ground-state charge-transfer complexes of anthocyanins with naturally occurring electron-donor copigments, such as hydroxylated flavones, flavonoids, and hydroxycinnamic or benzoic acids, do exhibit high UV-B absorptivities that complement that of the anthocyanins. In this work, we report a comparative study of the photophysics of the naturally occurring anthocyanin cyanin, intermol. cyanin-coumaric acid complexes, and an acylated anthocyanin, i.e., cyanin with a pendant coumaric ester copigment. Both inter- and intramol. anthocyanin-copigment complexes are shown to have ultrafast energy dissipation pathways comparable to those of model flavylium cation-copigment complexes. However, from the standpoint of photoprotection, the results indicate that the covalent attachment of copigment mols. to the anthocyanin represents a much more efficient strategy by providing the plant with significant UV-B absorption capacity and at the same time coupling this absorption to efficient energy dissipation pathways (ultrafast internal conversion of the complexed form and fast energy transfer from the excited copigment to the anthocyanin followed by adiabatic proton transfer) that avoid net photochem. damage.
- 163Song, B. J.; Sapper, T. N.; Burtch, C. E.; Brimmer, K.; Goldschmidt, M.; Ferruzzi, M. G. Photo- and Thermodegradation of Anthocyanins from Grape and Purple Sweet Potato in Model Beverage Systems J. Agric. Food Chem. 2013, 61 (6) 1364– 1372 DOI: 10.1021/jf3044007Google ScholarThere is no corresponding record for this reference.
- 164Malien-Aubert, C.; Dangles, O.; Amiot, M. J. Color Stability of Commercial Anthocyanin-Based Extracts in Relation to the Phenolic Composition. Protective Effects by Intra- and Intermolecular Copigmentation J. Agric. Food Chem. 2001, 49 (1) 170– 176 DOI: 10.1021/jf000791oGoogle ScholarThere is no corresponding record for this reference.
- 165Cabrita, L.; Petrov, V.; Pina, F. On the Thermal Degradation of Anthocyanidins: Cyanidin RSC Adv. 2014, 4 (36) 18939 DOI: 10.1039/c3ra47809bGoogle Scholar165On the thermal degradation of anthocyanidins: cyanidinCabrita, Luis; Petrov, Vesselin; Pina, FernandoRSC Advances (2014), 4 (36), 18939-18944CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Cyanidin was studied by direct pH jumps (from equilibrated solns. at very low pH values to higher pH values) and reverse pH jumps (from equilibrated or not equilibrated solns. at higher pH values to very low ones). The kinetic steps of the direct and reverse pH jumps were followed by stopped flow, absorption spectroscopy and HPLC, at different timescales. The pH dependent rate const. of the slower kinetic process to reach the equil. follows a bell shaped curve as described for many synthetic flavylium compds. Unlike anthocyanins, it was proved that there is no pH dependent reversibility in the system, since the chalcone suffers an irreversible degrdn. process. The math. expression to describe the bell shaped behavior was deduced. These results contribute to explain why in plants glycosylation is crucial for the stabilization of the anthocyanins.
- 166Schneider, H.-J. Dispersive Interactions in Solution Complexes Acc. Chem. Res. 2015, 48 (7) 1815– 1822 DOI: 10.1021/acs.accounts.5b00111Google Scholar166Dispersive Interactions in Solution ComplexesSchneider, Hans-JoergAccounts of Chemical Research (2015), 48 (7), 1815-1822CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Dispersive interactions are known to play a major role in mol. assocns. in the gas phase and in the solid state. In soln., however, their significance has been disputed in recent years on the basis of several arguments. A major problem until now has been the sepn. of dispersive and hydrophobic effects, which are both maximized in water due the low polarizability of this most important medium. Analyses of complexes between porphyrins and systematically varied substrates in water have allowed us to discriminate dispersive from hydrophobic effects, as the latter turned out to be negligible for complexations with flat surfaces such as porphyrins. Also, for the first time, it has become possible to obtain binding free energy increments ΔΔG for a multitude of org. residues including halogen, amide, amino, ether, carbonyl, ester, nitro, sulfur, unsatured, and cyclopropane groups, which turned out to be additive. Binding contributions for satd. residues are unmeasurably small, with ΔΔG > 1 kJ/mol, but they increase to, e.g., ΔΔG = 5 kJ/mol for a nitro group, a value not far from, e.g., that of a stacking pyridine ring. Stacking interactions of heteroarenes with porphyrins depend essentially on the size of the arenes, in line with polarizabilities, and seem to be rather independent of the position of nitrogen within the rings. Measurements of halogen derivs. indicate that complexes with porphyrins, cyclodextrins, and pillarenes as hosts in different media consistently show increasing stability from fluorine to iodine as the substituent. This, and the obsd. sequence with other substrates, is in line with the expected increase in dispersive forces with increasing polarizability. Induced dipoles, which also would increase with polarizability, can be ruled out as providing the driving source in view of the data with halides: the obsd. stability sequence is opposite the change of electronegativity from fluorine to iodine. The same holds for the solvent effect obsd. in ethanol-water mixts. Dispersive contributions vary not only with the polarizability of the used media but also with the interacting receptor sites; it has been shown that for cucurbiturils the polarizability inside the cavity is extremely low, which also explains why hydrophobic effects are maximized with these hosts. Complexations with other known host compds., however, such as those between cryptands or cavitands with, e.g., noble gases, bear the signature of dominating dispersive forces. Some recent examples illustrate that such van der Waals forces can also play an important role in complexations with proteins. Again, a clue for this is the increase in ΔG for inhibitor binding by 7 kJ/mol for, e.g., a bromine in comparison to a fluorine deriv.
- 167Hunter, C. A.; Sanders, J. K. M. The Nature of.pi.-.pi. Interactions J. Am. Chem. Soc. 1990, 112 (14) 5525– 5534 DOI: 10.1021/ja00170a016Google Scholar167The nature of π-π interactionsHunter, Christopher A.; Sanders, Jeremy K. M.Journal of the American Chemical Society (1990), 112 (14), 5525-34CODEN: JACSAT; ISSN:0002-7863.A simple model of the charge distribution in a π-system is used to explain the strong geometrical requirements for interactions between arom. mols. The key feature of the model is that it considers the σ-framework and the π-electrons sep. and demonstrates that net favorable π-π interactions are actually the result of π-σ attractions that overcome π-π repulsions. The calcns. correlate with observations made on porphyrin π-π interactions both in soln. and in the cryst. state. By using an idealized π-atom, some general rules for predicting the geometry of favorable π-π interactions are derived. In particular a favorable offset or slipped geometry is predicted. These rules successfully predict the geometry of intermol. interactions in the crystal structures of arom. mols. and rationalize a range of host-guest phenomena. The theory demonstrates that the electron donor-acceptor concept can be misleading: it is the properties of the atoms at the points of intermol. contact rather than the overall mol. properties which are important.
- 168Hwang, J.; Li, P.; Carroll, W. R.; Smith, M. D.; Pellechia, P. J.; Shimizu, K. D. Additivity of Substituent Effects in Aromatic Stacking Interactions J. Am. Chem. Soc. 2014, 136 (40) 14060– 14067 DOI: 10.1021/ja504378pGoogle Scholar168Additivity of Substituent Effects in Aromatic Stacking InteractionsHwang, Jungwun; Li, Ping; Carroll, William R.; Smith, Mark D.; Pellechia, Perry J.; Shimizu, Ken D.Journal of the American Chemical Society (2014), 136 (40), 14060-14067CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The goal of this study was to exptl. test the additivity of the electrostatic substituent effects (SEs) for the arom. stacking interaction. The additivity of the SEs was assessed using a small mol. model system that could adopt an offset face-to-face arom. stacking geometry. The intramol. interactions of these mol. torsional balances were quant. measured via the changes in a folded/unfolded conformational equil. Five different types of substituents were examd. (CH3, OCH3, Cl, CN, and NO2) that ranged from electron-donating to electron-withdrawing. The strength of the intramol. stacking interactions was measured for 21 substituted arom. stacking balances and 21 control balances in chloroform soln. The obsd. stability trends were consistent with additive SEs. Specifically, additive SE models could predict SEs with an accuracy from ±0.01 to ±0.02 kcal/mol. The additive SEs were consistent with Wheeler and Houk's direct SE model. However, the indirect or polarization SE model cannot be ruled out as it shows similar levels of additivity for two to three substituent systems, which were the no. of substituents in our model system. SE additivity also has practical utility as the SEs can be accurately predicted. This should aid in the rational design and optimization of systems that utilize arom. stacking interactions.
- 169Cockroft, S. L.; Hunter, C. A.; Lawson, K. R.; Perkins, J.; Urch, C. J. Electrostatic Control of Aromatic Stacking Interactions J. Am. Chem. Soc. 2005, 127 (24) 8594– 8595 DOI: 10.1021/ja050880nGoogle Scholar169Electrostatic Control of Aromatic Stacking InteractionsCockroft, Scott L.; Hunter, Christopher A.; Lawson, Kevin R.; Perkins, Julie; Urch, Christopher J.Journal of the American Chemical Society (2005), 127 (24), 8594-8595CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A supramol. approach has been used to investigate the free energies of intermol. arom. stacking interactions. Chem. double mutant cycles have been used to measure the effect of a range of substituents on face-to-face stacking interactions with Ph and pentafluorophenyl rings. Electrostatic effects dominate the trends in interaction energy.
- 170Meyer, E. A.; Castellano, R. K.; Diederich, F. Interactions with Aromatic Rings in Chemical and Biological Recognition Angew. Chem., Int. Ed. 2003, 42 (11) 1210– 1250 DOI: 10.1002/anie.200390319Google Scholar170Interactions with aromatic rings in chemical and biological recognitionMeyer, Emmanuel A.; Castellano, Ronald K.; Diederich, FrancoisAngewandte Chemie, International Edition (2003), 42 (11), 1210-1250CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Intermol. interactions involving arom. rings are key processes in both chem. and biol. recognition. Their understanding is essential for rational drug design and lead optimization in medicinal chem. Different approaches-biol. studies, mol. recognition studies with artificial receptors, crystallog. database mining, gas-phase studies, and theor. calcns. - are pursued to generate a profound understanding of the structural and energetic parameters of individual recognition modes involving arom. rings. This review attempts to combine and summarize the knowledge gained from these investigations. The review focuses mainly on examples with biol. relevance since one of its aims it to enhance the knowledge of mol. recognition forces that is essential for drug development.
- 171Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic Rings in Chemical and Biological Recognition: Energetics and Structures Angew. Chem., Int. Ed. 2011, 50 (21) 4808– 4842 DOI: 10.1002/anie.201007560Google Scholar171Aromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.
- 172Hunter, C. A. Quantifying Intermolecular Interactions: Guidelines for the Molecular Recognition Toolbox Angew. Chem., Int. Ed. 2004, 43 (40) 5310– 5324 DOI: 10.1002/anie.200301739Google Scholar172Quantifying intermolecular interactions: Guidelines for the molecular recognition toolboxHunter, Christopher A.Angewandte Chemie, International Edition (2004), 43 (40), 5310-5324CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Mol. recognition events in soln. are affected by many different factors that have hampered the development of an understanding of intermol. interactions at a quant. level. Tendency is to partition these effects into discrete phenomenol. fields that are classified, named, and divorced: arom. interactions, cation-T interactions, CH-O hydrogen bonds, short strong hydrogen bonds, and hydrophobic interactions to name a few. To progress in the field, the authors need to develop an integrated quant. appreciation of the relative magnitudes of all of the different effects that might influence the mol. recognition behavior of a given system. In an effort to navigate undergraduates through the vast and sometimes contradictory literature on the subject, I have developed an approach that treats theor. ideas and exptl. observations about intermol. interactions in the gas phase, the solid state, and soln. from a single simplistic viewpoint. The key features are outlined here, and although many of the ideas will be familiar, the aim is to provide a semiquant. thermodn. ranking of these effects in soln. at room temp.
- 173Trouillas, P.; Di Meo, F.; Gierschner, J.; Linares, M.; Sancho-García, J. C.; Otyepka, M. Optical Properties of Wine Pigments: Theoretical Guidelines with New Methodological Perspectives Tetrahedron 2015, 71 (20) 3079– 3088 DOI: 10.1016/j.tet.2014.10.046Google ScholarThere is no corresponding record for this reference.
- 174London, F. The General Theory of Molecular Forces Trans. Faraday Soc. 1937, 33 (0) 8b– 26 DOI: 10.1039/tf937330008bGoogle ScholarThere is no corresponding record for this reference.
- 175Chandler, D. Interfaces and the Driving Force of Hydrophobic Assembly Nature 2005, 437 (7059) 640– 647 DOI: 10.1038/nature04162Google Scholar175Interfaces and the driving force of hydrophobic assemblyChandler, DavidNature (London, United Kingdom) (2005), 437 (7059), 640-647CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. The hydrophobic effect - the tendency for oil and water to segregate - is important in diverse phenomena, from the cleaning of laundry, to the creation of micro-emulsions to make new materials, to the assembly of proteins into functional complexes. This effect is multifaceted depending on whether hydrophobic mols. are individually hydrated or driven to assemble into larger structures. Despite the basic principles underlying the hydrophobic effect being qual. well understood, only recently have theor. developments begun to explain and quantify many features of this ubiquitous phenomenon.
- 176Biedermann, F.; Nau, W. M.; Schneider, H.-J. The Hydrophobic Effect Revisited—Studies with Supramolecular Complexes Imply High-Energy Water as a Noncovalent Driving Force Angew. Chem., Int. Ed. 2014, 53 (42) 11158– 11171 DOI: 10.1002/anie.201310958Google Scholar176The Hydrophobic Effect Revisited-Studies with Supramolecular Complexes Imply High-Energy Water as a Noncovalent Driving ForceBiedermann, Frank; Nau, Werner M.; Schneider, Hans-JoergAngewandte Chemie, International Edition (2014), 53 (42), 11158-11171CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Traditional descriptions of the hydrophobic effect from entropic arguments or the calcn. of solvent-occupied surfaces must be questioned in view of new results obtained with supramol. complexes. In these studies, it was possible to sep. hydrophobic from dispersive interactions, which are strongest in aq. systems. Even very hydrophobic alkanes assoc. significantly only in cavities contg. water mols. with an insufficient no. of possible hydrogen bonds. The replacement of high-energy water in cavities by guest mols. is the essential enthalpic driving force for complexation, as borne out by data for complexes of cyclodextrins, cyclophanes, and cucurbiturils, for which complexation enthalpies of up to -100 kJ mol-1 were reached for encapsulated alkyl residues. Water-box simulations were used to characterize the different contributions from high-energy water and enabled the calcn. of the assocn. free enthalpies for selected cucurbituril complexes to within a 10% deviation from exptl. values. Cavities in artificial receptors are more apt to show the enthalpic effect of high-energy water than those in proteins or nucleic acids, because they bear fewer or no functional groups in the inner cavity to stabilize interior water mols.
- 177Waters, M. L. Aromatic Interactions Acc. Chem. Res. 2013, 46 (4) 873– 873 DOI: 10.1021/ar4000828Google Scholar177Aromatic InteractionsWaters, Marcey L.Accounts of Chemical Research (2013), 46 (4), 873CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)There is no expanded citation for this reference.
- 178Yang, L.; Adam, C.; Nichol, G. S.; Cockroft, S. L. How Much Do van Der Waals Dispersion Forces Contribute to Molecular Recognition in Solution? Nat. Chem. 2013, 5 (12) 1006– 1010 DOI: 10.1038/nchem.1779Google Scholar178How much do van der Waals dispersion forces contribute to molecular recognition in solution?Yang, Lixu; Adam, Catherine; Nichol, Gary S.; Cockroft, Scott L.Nature Chemistry (2013), 5 (12), 1006-1010CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)The emergent properties that arise from self-assembly and mol. recognition phenomena are a direct consequence of noncovalent interactions. Gas-phase measurements and computational methods point to the dominance of dispersion forces in mol. assocn., but solvent effects complicate the unambiguous quantification of these forces in soln. Here, the authors used synthetic mol. balances to measure interactions between apolar alkyl chains in 31 org., fluorous and aq. solvent environments. The exptl. interaction energies are an order of magnitude smaller than ests. of dispersion forces between alkyl chains that were derived from vaporization enthalpies and dispersion-cor. calcns. Instead, cohesive solvent-solvent interactions are the major driving force behind apolar assocn. in soln. Probably theor. models that implicate important roles for dispersion forces in mol. recognition events should be interpreted with caution in solvent-accessible systems.
- 179Mackerell, A. D. Empirical Force Fields for Biological Macromolecules: Overview and Issues J. Comput. Chem. 2004, 25 (13) 1584– 1604 DOI: 10.1002/jcc.20082Google Scholar179Empirical force fields for biological macromolecules: Overview and issuesMacKerell, Alexander D., Jr.Journal of Computational Chemistry (2004), 25 (13), 1584-1604CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A review. Empirical force field-based studies of biol. macromols. are becoming a common tool for investigating their structure-activity relationships at an at. level of detail. Such studies facilitate interpretation of exptl. data and allow for information not readily accessible to exptl. methods to be obtained. A large part of the success of empirical force field-based methods is the quality of the force fields combined with the algorithmic advances that allow for more accurate reprodn. of exptl. observables. Presented is an overview of the issues assocd. with the development and application of empirical force fields to biomol. systems. This is followed by a summary of the force fields commonly applied to the different classes of biomols.; proteins, nucleic acids, lipids, and carbohydrates. In addn., issues assocd. with computational studies on "heterogeneous" biomol. systems and the transferability of force fields to a wide range of org. mols. of pharmacol. interest are discussed.
- 180Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field J. Comput. Chem. 2004, 25 (9) 1157– 1174 DOI: 10.1002/jcc.20035Google Scholar180Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 181Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic Molecules J. Am. Chem. Soc. 1996, 118 (9) 2309– 2309 DOI: 10.1021/ja955032eGoogle Scholar181A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic Molecules [ Erratum for J. Am. Chem. Soc. 1995, 117, 5179-5197 ]Cornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M. Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1996), 118 (9), 2309CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)There is no expanded citation for this reference.
- 182Zgarbová, M.; Otyepka, M.; Sponer, J.; Hobza, P.; Jurecka, P. Large-Scale Compensation of Errors in Pairwise-Additive Empirical Force Fields: Comparison of AMBER Intermolecular Terms with Rigorous DFT-SAPT Calculations Phys. Chem. Chem. Phys. 2010, 12 (35) 10476– 10493 DOI: 10.1039/c002656eGoogle ScholarThere is no corresponding record for this reference.
- 183Šponer, J.; Banáš, P.; Jurečka, P.; Zgarbová, M.; Kührová, P.; Havrila, M.; Krepl, M.; Stadlbauer, P.; Otyepka, M. Molecular Dynamics Simulations of Nucleic Acids. From Tetranucleotides to the Ribosome J. Phys. Chem. Lett. 2014, 5 (10) 1771– 1782 DOI: 10.1021/jz500557yGoogle Scholar183Molecular Dynamics Simulations of Nucleic Acids. From Tetranucleotides to the RibosomeSponer, Jiri; Banas, Pavel; Jurecka, Petr; Zgarbova, Marie; Kuhrova, Petra; Havrila, Marek; Krepl, Miroslav; Stadlbauer, Petr; Otyepka, MichalJournal of Physical Chemistry Letters (2014), 5 (10), 1771-1782CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)A review. We present a brief overview of explicit solvent mol. dynamics (MD) simulations of nucleic acids. We explain phys. chem. limitations of the simulations, namely, the mol. mechanics (MM) force field (FF) approxn. and limited time scale. Further, we discuss relations and differences between simulations and expts., compare std. and enhanced sampling simulations, discuss the role of starting structures, comment on different versions of nucleic acid FFs, and relate MM computations with contemporary quantum chem. Despite its limitations, we show that MD is a powerful technique for studying the structural dynamics of nucleic acids with a fast growing potential that substantially complements exptl. results and aids their interpretation.
- 184Adcock, S. A.; McCammon, J. A. Molecular Dynamics: Survey of Methods for Simulating the Activity of Proteins Chem. Rev. 2006, 106 (5) 1589– 1615 DOI: 10.1021/cr040426mGoogle Scholar184Molecular dynamics: Survey of methods for simulating the activity of proteinsAdcock, Stewart A.; McCammon, J. AndrewChemical Reviews (Washington, DC, United States) (2006), 106 (5), 1589-1615CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Mol. dynamics simulations (MDS) of proteins have provided many insights into the internal motions of these biomols. Simulation of in silico models aids in the interpretation and reconciliation of exptl. data. With ongoing advances in both methodol. and computational resources, MDS are being extended to larger systems and longer time scales. This enables the investigation of motions and conformational changes that have functional implications and yields information that is not available though any other means. Today's results suggest that (subject to the continuing utilization of synergies between expt. and simulation) the applications of MDS will command an increasingly crit. role in the understanding of biol. systems.
- 185Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. The Amber Biomolecular Simulation Programs J. Comput. Chem. 2005, 26 (16) 1668– 1688 DOI: 10.1002/jcc.20290Google Scholar185The amber biomolecular simulation programsCase, David A.; Cheatham, Thomas E., III; Darden, Tom; Gohlke, Holger; Luo, Ray; Merz, Kenneth M., Jr.; Onufriev, Alexey; Simmerling, Carlos; Wang, Bing; Woods, Robert J.Journal of Computational Chemistry (2005), 26 (16), 1668-1688CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The authors describe the development, current features, and some directions for future development of the Amber package of computer programs. This package evolved from a program that was constructed in the late 1970s to do Assisted Model Building with Energy Refinement, and now contains a group of programs embodying a no. of powerful tools of modern computational chem., focused on mol. dynamics and free energy calcns. of proteins, nucleic acids, and carbohydrates.
- 186Brooks, B. R.; Brooks, C. L.; MacKerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S. CHARMM: The Biomolecular Simulation Program J. Comput. Chem. 2009, 30 (10) 1545– 1614 DOI: 10.1002/jcc.21287Google Scholar186CHARMM: The biomolecular simulation programBrooks, B. R.; Brooks, C. L., III; Mackerell, A. D., Jr.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; Yang, W.; York, D. M.; Karplus, M.Journal of Computational Chemistry (2009), 30 (10), 1545-1614CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A review. CHARMM (Chem. at HARvard Mol. Mechanics) is a highly versatile and widely used mol. simulation program. It has been developed over the last three decades with a primary focus on mols. of biol. interest, including proteins, peptides, lipids, nucleic acids, carbohydrates, and small mol. ligands, as they occur in soln., crystals, and membrane environments. For the study of such systems, the program provides a large suite of computational tools that include numerous conformational and path sampling methods, free energy estimators, mol. minimization, dynamics, and anal. techniques, and model-building capabilities. The CHARMM program is applicable to problems involving a much broader class of many-particle systems. Calcns. with CHARMM can be performed using a no. of different energy functions and models, from mixed quantum mech.-mol. mech. force fields, to all-atom classical potential energy functions with explicit solvent and various boundary conditions, to implicit solvent and membrane models. The program has been ported to numerous platforms in both serial and parallel architectures. This article provides an overview of the program as it exists today with an emphasis on developments since the publication of the original CHARMM article in 1983. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009.
- 187Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD J. Comput. Chem. 2005, 26 (16) 1781– 1802 DOI: 10.1002/jcc.20289Google Scholar187Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 188Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C. GROMACS: Fast, Flexible, and Free J. Comput. Chem. 2005, 26 (16) 1701– 1718 DOI: 10.1002/jcc.20291Google Scholar188GROMACS: Fast, flexible, and freeVan Der Spoel, David; Lindahl, Erik; Hess, Berk; Groenhof, Gerrit; Mark, Alan E.; Berendsen, Herman J. C.Journal of Computational Chemistry (2005), 26 (16), 1701-1718CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)This article describes the software suite GROMACS (Groningen MAchine for Chem. Simulation) that was developed at the University of Groningen, The Netherlands, in the early 1990s. The software, written in ANSI C, originates from a parallel hardware project, and is well suited for parallelization on processor clusters. By careful optimization of neighbor searching and of inner loop performance, GROMACS is a very fast program for mol. dynamics simulation. It does not have a force field of its own, but is compatible with GROMOS, OPLS, AMBER, and ENCAD force fields. In addn., it can handle polarizable shell models and flexible constraints. The program is versatile, as force routines can be added by the user, tabulated functions can be specified, and analyses can be easily customized. Nonequil. dynamics and free energy detns. are incorporated. Interfaces with popular quantum-chem. packages (MOPAC, GAMES-UK, GAUSSIAN) are provided to perform mixed MM/QM simulations. The package includes about 100 utility and anal. programs. GROMACS is in the public domain and distributed (with source code and documentation) under the GNU General Public License. It is maintained by a group of developers from the Universities of Groningen, Uppsala, and Stockholm, and the Max Planck Institute for Polymer Research in Mainz. Its Web site is http://www.gromacs.org.
- 189Leach, A. R. Molecular Modelling: Principles and Applications; Pearson, 2001.Google ScholarThere is no corresponding record for this reference.
- 190Vashisth, H.; Skiniotis, G.; Brooks, C. L. Collective Variable Approaches for Single Molecule Flexible Fitting and Enhanced Sampling Chem. Rev. 2014, 114 (6) 3353– 3365 DOI: 10.1021/cr4005988Google ScholarThere is no corresponding record for this reference.
- 191Bruccoleri, R. E.; Karplus, M. Conformational Sampling Using High-Temperature Molecular Dynamics Biopolymers 1990, 29 (14) 1847– 1862 DOI: 10.1002/bip.360291415Google ScholarThere is no corresponding record for this reference.
- 192Chocholoušová, J.; Vacek, J.; Hobza, P. Acetic Acid Dimer in the Gas Phase, Nonpolar Solvent, Microhydrated Environment, and Dilute and Concentrated Acetic Acid: Ab Initio Quantum Chemical and Molecular Dynamics Simulations J. Phys. Chem. A 2003, 107 (17) 3086– 3092 DOI: 10.1021/jp027637kGoogle Scholar192Acetic Acid Dimer in the Gas Phase, Nonpolar Solvent, Microhydrated Environment, and Dilute and Concentrated Acetic Acid: Ab Initio Quantum Chemical and Molecular Dynamics SimulationsChocholousova, Jana; Vacek, Jaroslav; Hobza, PavelJournal of Physical Chemistry A (2003), 107 (17), 3086-3092CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Theor. study of the acetic acid dimer, its microhydration and its behavior in water and chloroform soln. was performed. To characterize the system, the authors adopted ab initio methods at the DFT and RI MP2 (the resoln. of the identity approxn. MP2) levels for the gas-phase calcns., PCM (polarizable continuum model) approxn. using the polarizable conductor calcn. model (COSMO) for description of solvent, and const. energy (NVE) and const. temp. (NVT) mol. dynamics simulations for gas phase and explicit solvent calcns., resp. The cyclic structure of the acetic acid dimer is the most stable in the gas phase only. During microhydration, the water mols. are incorporated in the dimer leading to water-sepd. structures. This conclusion is based on ab initio quantum chem. calcns., as well as on mol. dynamics simulations. The fact that the cyclic structure does not appear in water soln. is in agreement with previous theor. and exptl. results. Extending the search also on other acetic acid dimer structures, acetic acid does not form any dimer structure in water soln. The cyclic structure also probably is stable in chloroform soln.
- 193Zelený, T.; Hobza, P.; Kabelác, M. Microhydration of Guanine···cytosine Base Pairs, a Theoretical Study on the Role of Water in Stability, Structure and Tautomeric Equilibrium Phys. Chem. Chem. Phys. 2009, 11 (18) 3430– 3435 DOI: 10.1039/b819350aGoogle ScholarThere is no corresponding record for this reference.
- 194Sugita, Y.; Okamoto, Y. Replica-Exchange Molecular Dynamics Method for Protein Folding Chem. Phys. Lett. 1999, 314 (1–2) 141– 151 DOI: 10.1016/S0009-2614(99)01123-9Google Scholar194Replica-exchange molecular dynamics method for protein foldingSugita, Y.; Okamoto, Y.Chemical Physics Letters (1999), 314 (1,2), 141-151CODEN: CHPLBC; ISSN:0009-2614. (Elsevier Science B.V.)We have developed a formulation for mol. dynamics algorithm for the replica-exchange method. The effectiveness of the method for the protein-folding problem is tested with the penta-peptide Met-enkephalin. The method can overcome the multiple-min. problem by exchanging non-interacting replicas of the system at several temps. From only one simulation run, one can obtain probability distributions in canonical ensemble for a wide temp. range using multiple-histogram re-weighting techniques, which allows the calcn. of any thermodn. quantity as a function of temp. in that range.
- 195Beck, D. A. C.; White, G. W. N.; Daggett, V. Exploring the Energy Landscape of Protein Folding Using Replica-Exchange and Conventional Molecular Dynamics Simulations J. Struct. Biol. 2007, 157 (3) 514– 523 DOI: 10.1016/j.jsb.2006.10.002Google ScholarThere is no corresponding record for this reference.
- 196Laio, A.; Parrinello, M. Escaping Free-Energy Minima Proc. Natl. Acad. Sci. U. S. A. 2002, 99 (20) 12562– 12566 DOI: 10.1073/pnas.202427399Google Scholar196Escaping free-energy minimaLaio, Alessandro; Parrinello, MicheleProceedings of the National Academy of Sciences of the United States of America (2002), 99 (20), 12562-12566CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)We introduce a powerful method for exploring the properties of the multidimensional free energy surfaces (FESs) of complex many-body systems by means of coarse-grained non-Markovian dynamics in the space defined by a few collective coordinates. A characteristic feature of these dynamics is the presence of a history-dependent potential term that, in time, fills the min. in the FES, allowing the efficient exploration and accurate detn. of the FES as a function of the collective coordinates. We demonstrate the usefulness of this approach in the case of the dissocn. of a NaCl mol. in water and in the study of the conformational changes of a dialanine in soln.
- 197Barducci, A.; Bonomi, M.; Parrinello, M. Metadynamics Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1 (5) 826– 843 DOI: 10.1002/wcms.31Google Scholar197MetadynamicsBarducci, Alessandro; Bonomi, Massimiliano; Parrinello, MicheleWiley Interdisciplinary Reviews: Computational Molecular Science (2011), 1 (5), 826-843CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Metadynamics is a powerful technique for enhancing sampling in mol. dynamics simulations and reconstructing the free-energy surface as a function of few selected degrees of freedom, often referred to as collective variables (CVs). In metadynamics, sampling is accelerated by a history-dependent bias potential, which is adaptively constructed in the space of the CVs. Since its first appearance, significant improvements have been made to the original algorithm, leading to an efficient, flexible, and accurate method that has found many successful applications in several domains of science. Here, we discuss first the theory underlying metadynamics and its recent developments. In particular, we focus on the crucial issue of choosing an appropriate set of CVs and on the possible strategies to alleviate this difficulty. Later in the second part, we present a few recent representative applications, which we have classified into three main classes: solid-state physics, chem. reactions, and biomols.
- 198Brás, N. F.; Cruz, L.; Fernandes, P. A.; De Freitas, V.; Ramos, M. J. Conformational Study of Two Diasteroisomers of Vinylcatechin Dimers in a Methanol Solution Int. J. Quantum Chem. 2011, 111 (7–8) 1498– 1510 DOI: 10.1002/qua.22631Google ScholarThere is no corresponding record for this reference.
- 199Kollman, P. A.; Massova, I.; Reyes, C.; Kuhn, B.; Huo, S.; Chong, L.; Lee, M.; Lee, T.; Duan, Y.; Wang, W. Calculating Structures and Free Energies of Complex Molecules: Combining Molecular Mechanics and Continuum Models Acc. Chem. Res. 2000, 33 (12) 889– 897 DOI: 10.1021/ar000033jGoogle Scholar199Calculating Structures and Free Energies of Complex Molecules: Combining Molecular Mechanics and Continuum ModelsKollman, Peter A.; Massova, Irina; Reyes, Carolina; Kuhn, Bernd; Huo, Shuanghong; Chong, Lillian; Lee, Matthew; Lee, Taisung; Duan, Yong; Wang, Wei; Donini, Oreola; Cieplak, Piotr; Srinivasan, Jaysharee; Case, David A.; Cheatham, Thomas E., IIIAccounts of Chemical Research (2000), 33 (12), 889-897CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review, with 63 refs. A historical perspective on the application of mol. dynamics (MD) to biol. macromols. is presented. Recent developments combining state-of-the-art force fields with continuum solvation calcns. have allowed us to reach the fourth era of MD applications in which one can often derive both accurate structure and accurate relative free energies from mol. dynamics trajectories. We illustrate such applications on nucleic acid duplexes, RNA hairpins, protein folding trajectories, and protein-ligand, protein-protein, and protein-nucleic acid interactions.
- 200Bahar, I.; Lezon, T. R.; Bakan, A.; Shrivastava, I. H. Normal Mode Analysis of Biomolecular Structures: Functional Mechanisms of Membrane Proteins Chem. Rev. 2010, 110 (3) 1463– 1497 DOI: 10.1021/cr900095eGoogle Scholar200Normal Mode Analysis of Biomolecular Structures: Functional Mechanisms of Membrane ProteinsBahar, Ivet; Lezon, Timothy R.; Bakan, Ahmet; Shrivastava, Indira H.Chemical Reviews (Washington, DC, United States) (2010), 110 (3), 1463-1497CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Structural dynamics and function of membrane proteins are discussed in relation to principal component anal. of exptl. resolved conformations, normal mode anal. and elastic network models.
- 201Andricioaei, I.; Karplus, M. On the Calculation of Entropy from Covariance Matrices of the Atomic Fluctuations J. Chem. Phys. 2001, 115 (14) 6289– 6292 DOI: 10.1063/1.1401821Google Scholar201On the calculation of entropy from covariance matrices of the atomic fluctuationsAndricioaei, Ioan; Karplus, MartinJournal of Chemical Physics (2001), 115 (14), 6289-6292CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)An ad hoc method for calcg. the entropy of a biomol. system from the covariance matrix of the at. fluctuations is analyzed. It is shown that its essential assumption can be eliminated by a quasiharmonic anal. The computer time required for use of the latter is of the same order as that of the former.
- 202Tsui, V.; Case, D. A. Theory and Applications of the Generalized Born Solvation Model in Macromolecular Simulations Biopolymers 2000, 56 (4) 275– 291 DOI: 10.1002/1097-0282(2000)56:4<275::AID-BIP10024>3.0.CO;2-EGoogle Scholar202Theory and applications of the generalized Born solvation model in macromolecular simulationsTsui V; Case D ABiopolymers (2000-2001), 56 (4), 275-91 ISSN:0006-3525.Generalized Born (GB) models provide an attractive way to include some thermodynamic aspects of aqueous solvation into simulations that do not explicitly model the solvent molecules. Here we discuss our recent experience with this model, presenting in detail the way it is implemented and parallelized in the AMBER molecular modeling code. We compare results using the GB model (or GB plus a surface-area based "hydrophobic" term) to explicit solvent simulations for a 10 base-pair DNA oligomer, and for the 108-residue protein thioredoxin. A slight modification of our earlier suggested parameters makes the GB results more like those found in explicit solvent, primarily by slightly increasing the strength of NH [bond] O and NH [bond] N internal hydrogen bonds. Timing and energy stability results are reported, with an eye toward using these model for simulations of larger macromolecular systems and longer time scales.
- 203Rocchia, W.; Alexov, E.; Honig, B. Extending the Applicability of the Nonlinear Poisson–Boltzmann Equation: Multiple Dielectric Constants and Multivalent Ions J. Phys. Chem. B 2001, 105 (28) 6507– 6514 DOI: 10.1021/jp010454yGoogle Scholar203Extending the applicability of the nonlinear Poisson-Boltzmann equation: multiple dielectric constants and multivalent ionsRocchia, W.; Alexov, E.; Honig, B.Journal of Physical Chemistry B (2001), 105 (28), 6507-6514CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new version of the DelPhi program, which provides numerical solns. to the nonlinear Poisson-Boltzmann (PB) equation, is reported. The program can divide space into multiple regions contg. different dielec. consts. and can treat systems contg. mixed salt solns. where the valence and concn. of each ion is different. The electrostatic free energy is calcd. by decompg. the various energy terms into Coulombic interactions so that that the calcd. free energies are independent of the lattice used to solve the PB equation. This, together with algorithms that optimally position polarization charges on the mol. surface, leads to a significant decrease in the dependence of the electrostatic free energy on the resoln. of the lattice used to solve the PB equation and, hence, to a remarkable improvement in the precision of the calcd. values. The Gauss-Seidel algorithm used in the current version of DelPhi is retained so that the new program retains many of the optimization features of the old one. The program uses dynamic memory allocation and can easily handle systems requiring large grid dimensions - for example, a 3003 system can be conveniently treated on a single SGI R12000 processor. An algorithm that ests. the best relaxation parameter to solve the nonlinear equation for a given system is described, and is implemented in the program at run time. A no. of applications of the program are presented.
- 204Baker, C. M.; Lopes, P. E. M.; Zhu, X.; Roux, B.; MacKerell, A. D. Accurate Calculation of Hydration Free Energies Using Pair-Specific Lennard-Jones Parameters in the CHARMM Drude Polarizable Force Field J. Chem. Theory Comput. 2010, 6 (4) 1181– 1198 DOI: 10.1021/ct9005773Google Scholar204Accurate Calculation of Hydration Free Energies using Pair-Specific Lennard-Jones Parameters in the CHARMM Drude Polarizable Force FieldBaker, Christopher M.; Lopes, Pedro E. M.; Zhu, Xiao; Roux, Benoit; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2010), 6 (4), 1181-1198CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Lennard-Jones (LJ) parameters for a variety of model compds. have previously been optimized within the CHARMM Drude polarizable force field to reproduce accurately pure liq. phase thermodn. properties as well as addnl. target data. While the polarizable force field resulting from this optimization procedure has been shown to satisfactorily reproduce a wide range of exptl. ref. data across numerous series of small mols., a slight but systematic overestimate of the hydration free energies has also been noted. Here, the reprodn. of exptl. hydration free energies is greatly improved by the introduction of pair-specific LJ parameters between solute heavy atoms and water oxygen atoms that override the std. LJ parameters obtained from combining rules. The changes are small and a systematic protocol is developed for the optimization of pair-specific LJ parameters and applied to the development of pair-specific LJ parameters for alkanes, alcs., and ethers. The resulting parameters not only yield hydration free energies in good agreement with exptl. values, but also provide a framework upon which other pair-specific LJ parameters can be added as new compds. are parametrized within the CHARMM Drude polarizable force field. Detailed anal. of the contributions to the hydration free energies reveals that the dispersion interaction is the main source of the systematic errors in the hydration free energies. This information suggests that the systematic error may result from problems with the LJ combining rules and is combined with anal. of the pair-specific LJ parameters obtained in this work to identify a preliminary improved combining rule.
- 205Allinger, N. L. Conformational Analysis. 130. MM2. A Hydrocarbon Force Field Utilizing V1 and V2 Torsional Terms J. Am. Chem. Soc. 1977, 99 (25) 8127– 8134 DOI: 10.1021/ja00467a001Google Scholar205Conformational analysis. 130. MM2. A hydrocarbon force field utilizing V1 and V2 torsional termsAllinger, Norman L.Journal of the American Chemical Society (1977), 99 (25), 8127-34CODEN: JACSAT; ISSN:0002-7863.An improved force field for mol.-mechanics calcns. of the structures and energies of hydrocarbons is presented. The problem of simultaneously obtaining a sufficiently large gauche butane interaction energy while keeping the H atoms small enough for good structural predictions was solved with 1- and 2-fold rotational barriers. For 42 selected diverse types of hydrocarbons, the std. deviation between the calcd. and exptl. heats of formation is 0.42 kcal/mol, compared with an av. reported exptl. error for the same group of compds. of 0.40 kcal/mol.
- 206Charlton, A. J.; Davis, A. L.; Jones, D. P.; Lewis, J. R.; Davies, A. P.; Haslam, E.; Williamson, M. P. The Self-Association of the Black Tea Polyphenol Theaflavin and Its Complexation with Caffeine J. Chem. Soc. Perkin Trans. 2 2000, 2) 317– 322 DOI: 10.1039/a906380cGoogle ScholarThere is no corresponding record for this reference.
- 207Kunsági-Máté, S.; Szabó, K.; Nikfardjam, M. P.; Kollár, L. Determination of the Thermodynamic Parameters of the Complex Formation between Malvidin-3-O-Glucoside and Polyphenols. Copigmentation Effect in Red Wines J. Biochem. Biophys. Methods 2006, 69 (1–2) 113– 119 DOI: 10.1016/j.jbbm.2006.03.014Google Scholar207Determination of the thermodynamic parameters of the complex formation between malvidin-3-O-glucoside and polyphenols. Copigmentation effect in red winesKunsagi-Mate, Sandor; Szabo, Kornelia; Nikfardjam, Martin P.; Kollar, LaszloJournal of Biochemical and Biophysical Methods (2006), 69 (1-2), 113-119CODEN: JBBMDG; ISSN:0165-022X. (Elsevier Ltd.)The thermodn. of the mol. assocn. process between the malvidin-3-O-glucoside and a series of polyphenol derivs. (called copigmentation' in food chem.) were studied in aq. media. The Gibbs free energy, enthalpy and entropy values were detd. by the fluorometric method. A combination of the Job's method with the van't Hoff theory was applied for data evaluation. The results show the exothermic character of the copigmentation process. The change of the enthalpy seems to be the same in every complexation step. However, the decreasing of the entropy term is higher at higher stoichiometries. As a result, the Gibbs free energy changes and, thus, the complex stability decreases quickly with increasing stoichiometry. Quantum-chem. investigation reveals the complexity of mol. interactions between malvidin and polyphenols, which is preferably based on π-π and OH-π interaction moderated by repulsive Coulomb-type interactions.
- 208Kalisz, S.; Oszmiański, J.; Hładyszowski, J.; Mitek, M. Stabilization of Anthocyanin and Skullcap Flavone Complexes – Investigations with Computer Simulation and Experimental Methods Food Chem. 2013, 138 (1) 491– 500 DOI: 10.1016/j.foodchem.2012.10.146Google ScholarThere is no corresponding record for this reference.
- 209Kunsági-Máté, S.; Ortmann, E.; Kollár, L.; Nikfardjam, M. P. Effect of the Solvatation Shell Exchange on the Formation of Malvidin- 3- O -Glucoside–Ellagic Acid Complexes J. Phys. Chem. B 2007, 111 (40) 11750– 11755 DOI: 10.1021/jp0740144Google ScholarThere is no corresponding record for this reference.
- 210Sousa, A.; Araújo, P.; Cruz, L.; Brás, N. F.; Mateus, N.; De Freitas, V. Evidence for Copigmentation Interactions between Deoxyanthocyanidin Derivatives (Oaklins) and Common Copigments in Wine Model Solutions J. Agric. Food Chem. 2014, 62 (29) 6995– 7001 DOI: 10.1021/jf404640mGoogle ScholarThere is no corresponding record for this reference.
- 211Košinová, P.; Berka, K.; Wykes, M.; Otyepka, M.; Trouillas, P. Positioning of Antioxidant Quercetin and Its Metabolites in Lipid Bilayer Membranes: Implication for Their Lipid-Peroxidation Inhibition J. Phys. Chem. B 2012, 116 (4) 1309– 1318 DOI: 10.1021/jp208731gGoogle Scholar211Positioning of Antioxidant Quercetin and Its Metabolites in Lipid Bilayer Membranes: Implication for Their Lipid-Peroxidation InhibitionKosinova, Pavlina; Berka, Karel; Wykes, Michael; Otyepka, Michal; Trouillas, PatrickJournal of Physical Chemistry B (2012), 116 (4), 1309-1318CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Among numerous biol. activities, natural polyphenols are antioxidants widely distributed in plants capable of inhibiting lipid peroxidn., which belongs to the most serious degenerative cell processes. Positioning of antioxidants in lipid bilayers can provide an insight to the lipid-peroxidn. inhibition at the mol. level. This work aims at detg. the location and orientation of quercetin and its most representative (glucuronidated, methylated, and sulfated) metabolites in lipid bilayer via mol. dynamic simulations. We show that quercetin derivs. penetrate the lipid bilayer and that the depths of penetration depend on mol. charge and substitutional variations. In the presence of charged substituents (sulfates and glucuronidates), the mol. is pulled toward the lipid bilayer surface. The orientation also depends on substitution as H-bonds are formed between the polar head groups of the bilayer and the (i) OH groups, (ii) sugar, and (iii) sulfate moieties of the antioxidants. As flavonoids and their derivs. are preferentially localized in the lipid bilayer membrane or on the bilayer/water interface, they readily conc. in a relatively narrow membrane region. Despite the low concns. of flavonoids in food, their spatial confinement in the membrane greatly enhances their local concn. in this vital region, thus increasing their importance for in vivo biol. activities including oxidative stress defense.
- 212Sirk, T. W.; Brown, E. F.; Sum, A. K.; Friedman, M. Molecular Dynamics Study on the Biophysical Interactions of Seven Green Tea Catechins with Lipid Bilayers of Cell Membranes J. Agric. Food Chem. 2008, 56 (17) 7750– 7758 DOI: 10.1021/jf8013298Google Scholar212Molecular Dynamics Study on the Biophysical Interactions of Seven Green Tea Catechins with Lipid Bilayers of Cell MembranesSirk, Timothy W.; Brown, Eugene F.; Sum, Amadeu K.; Friedman, MendelJournal of Agricultural and Food Chemistry (2008), 56 (17), 7750-7758CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Mol. dynamics simulations were performed to study the interactions of bioactive catechins (flavonoids) commonly found in green tea with lipid bilayers, as a model for cell membranes. Previously, multiple exptl. studies rationalized catechin's anticarcinogenic, antibacterial, and other beneficial effects in terms of physicochem. mol. interactions with the cell membranes. To contribute toward understanding the mol. role of catechins on the structure of cell membranes, we present simulation results for seven green tea catechins in lipid bilayer systems representative of HepG2 cancer cells. Our simulations show that the seven tea catechins evaluated have a strong affinity for the lipid bilayer via hydrogen bonding to the bilayer surface, with some of the smaller catechins able to penetrate underneath the surface. Epigallocatechin-gallate (EGCG) showed the strongest interaction with the lipid bilayer based on the no. of hydrogen bonds formed with lipid headgroups. The simulations also provide insight into the functional characteristics of the catechins that distinguish them as effective compds. to potentially alter the lipid bilayer properties. The results on the hydrogen-bonding effects, described here for the first time, may contribute to a better understanding of proposed multiple mol. mechanisms of the action of catechins in microorganisms, cancer cells, and tissues.
- 213Paloncýová, M.; Fabre, G.; DeVane, R. H.; Trouillas, P.; Berka, K.; Otyepka, M. Benchmarking of Force Fields for Molecule–Membrane Interactions J. Chem. Theory Comput. 2014, 10 (9) 4143– 4151 DOI: 10.1021/ct500419bGoogle Scholar213Benchmarking of Force Fields for Molecule-Membrane InteractionsPaloncyova, Marketa; Fabre, Gabin; DeVane, Russell H.; Trouillas, Patrick; Berka, Karel; Otyepka, MichalJournal of Chemical Theory and Computation (2014), 10 (9), 4143-4151CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Free energy profiles of eleven compds. with a model dimyristoylphosphatidylcholine (DMPC) membrane bilayer have been calcd. using five force fields, namely Berger, Slipids, CHARMM36, GAFFlipids, and GROMOS 43A1-S3 via mol. dynamics simulations. For the sake of comparison, the semicontinuous tool COSMOmic was also used. High correlation was obsd. between theor. and exptl. partition coeffs. (log K). Partition coeffs. calcd. by all-at. force fields (Slipids, CHARMM36, and GAFFlipids) and COSMOmic differed by less than 0.75 log units from the expt. and Slipids emerged as the best performing force field. This work provides the following recommendations (i) for a global, systematic and high throughput thermodn. evaluations (e.g., log K) of drugs COSMOmic is a tool of choice due to low computational costs; (ii) for studies of the hydrophilic mols. CHARMM36 should be considered; and (iii) for studies of more complex systems, taking into account all pros and cons, Slipids is the force field of choice.
- 214Podloucká, P.; Berka, K.; Fabre, G.; Paloncýová, M.; Duroux, J.-L.; Otyepka, M.; Trouillas, P. Lipid Bilayer Membrane Affinity Rationalizes Inhibition of Lipid Peroxidation by a Natural Lignan Antioxidant J. Phys. Chem. B 2013, 117 (17) 5043– 5049 DOI: 10.1021/jp3127829Google ScholarThere is no corresponding record for this reference.
- 215Zhou, H.-X.; Gilson, M. K. Theory of Free Energy and Entropy in Noncovalent Binding Chem. Rev. 2009, 109 (9) 4092– 4107 DOI: 10.1021/cr800551wGoogle Scholar215Theory of Free Energy and Entropy in Noncovalent BindingZhou, Huan-Xiang; Gilson, Michael K.Chemical Reviews (Washington, DC, United States) (2009), 109 (9), 4092-4107CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Theory of free energy and entropy in noncovalent binding aims to support the development of well-founded models of binding and meaningful interpretation of exptl. data. Discussion includes the statistical thermodn. of binding., changes in translational and other entropy components on binding.
- 216Černỳ, J.; Hobza, P. Non-Covalent Interactions in Biomacromolecules Phys. Chem. Chem. Phys. 2007, 9 (39) 5291– 5303 DOI: 10.1039/b704781aGoogle ScholarThere is no corresponding record for this reference.
- 217Grimme, S.; Antony, J.; Schwabe, T.; Mück-Lichtenfeld, C. Density Functional Theory with Dispersion Corrections for Supramolecular Structures, Aggregates, and Complexes of (bio) Organic Molecules Org. Biomol. Chem. 2007, 5 (5) 741– 758 DOI: 10.1039/b615319bGoogle ScholarThere is no corresponding record for this reference.
- 218Sherrill, C. D. Computations of Noncovalent N Interactions Rev. Comput. Chem. 2008, 26, 1 DOI: 10.1002/9780470399545.ch1Google ScholarThere is no corresponding record for this reference.
- 219Tschumper, G. S. Computations for Weak Noncovalent Rev. Comput. Chem. 2008, 26, 39 DOI: 10.1002/9780470399545.ch2Google ScholarThere is no corresponding record for this reference.
- 220Foster, M. E.; Sohlberg, K. Empirically Corrected DFT and Semi-Empirical Methods for Non-Bonding Interactions Phys. Chem. Chem. Phys. 2010, 12 (2) 307– 322 DOI: 10.1039/B912859JGoogle ScholarThere is no corresponding record for this reference.
- 221Riley, K. E.; Pitoňák, M.; Jurečka, P.; Hobza, P. Stabilization and Structure Calculations for Noncovalent Interactions in Extended Molecular Systems Based on Wave Function and Density Functional Theories Chem. Rev. 2010, 110 (9) 5023– 5063 DOI: 10.1021/cr1000173Google Scholar221Stabilization and Structure Calculations for Noncovalent Interactions in Extended Molecular Systems Based on Wave Function and Density Functional TheoriesRiley, Kevin E.; Pitonak, Michal; Jurecka, Petr; Hobza, PavelChemical Reviews (Washington, DC, United States) (2010), 110 (9), 5023-5063CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Topics covered include: wave function theory; MP2; methods from MP3 to CCSD; d. functional theories; empirical dispersion corrections; plane wave codes and nonlocal pseudopotentials; symmetry-adapted perturbation theory; and semiempirical quantum chem. theories.
- 222Hobza, P. The Calculation of Intermolecular Interaction Energies Annu. Rep. Prog. Chem., Sect. C: Phys. Chem. 2011, 107, 148– 168 DOI: 10.1039/c1pc90005fGoogle Scholar222The calculation of intermolecular interaction energiesHobza, PavelAnnual Reports on the Progress of Chemistry, Section C: Physical Chemistry (2011), 107 (), 148-168CODEN: ACPCDW; ISSN:0260-1826. (Royal Society of Chemistry)All life on our earth can be viewed as an application of supramol. chem., with noncovalent interactions playing a central role. The knowledge of total interaction energies as well as their components is topical for understanding the nature of these interactions and, in a broader sense, also for understanding the nature of stabilization of noncovalent systems like biomacromols. Accurate data on interaction energies can only be obtained from coupled-cluster with single and double and perturbative triple excitations (CCSD(T)) calcns. performed with extended basis sets. The CCSDD(T) calcns. thus provide benchmark data which can be used for testing and/or parameterizing other, computationally economical techniques. In the present review the applicability and performance of various recently introduced wavefunction and d. functional methods are examd. in detail.
- 223Grimme, S. Density Functional Theory with London Dispersion Corrections Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1 (2) 211– 228 DOI: 10.1002/wcms.30Google Scholar223Density functional theory with london dispersion correctionsGrimme, StefanWiley Interdisciplinary Reviews: Computational Molecular Science (2011), 1 (2), 211-228CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Dispersion corrections to std. Kohn-Sham d. functional theory (DFT) are reviewed. The focus is on computationally efficient methods for large systems that do not depend on virtual orbitals or rely on sepd. fragments. The recommended approaches (van der Waals d. functional and DFT-D) are asymptotically correct and can be used in combination with std. or slightly modified (short-range) exchange-correlation functionals. The importance of the dispersion energy in intramol. cases (conformational problems and thermochem.) is highlighted.
- 224Klimeš, J.; Michaelides, A. Perspective: Advances and Challenges in Treating van Der Waals Dispersion Forces in Density Functional Theory J. Chem. Phys. 2012, 137 (12) 120901 DOI: 10.1063/1.4754130Google Scholar224Perspective: Advances and challenges in treating van der Waals dispersion forces in density functional theoryKlimes, Jiri; Michaelides, AngelosJournal of Chemical Physics (2012), 137 (12), 120901/1-120901/12CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A review. Electron dispersion forces play a crucial role in detg. the structure and properties of biomols., mol. crystals, and many other systems. However, an accurate description of dispersion is highly challenging, with the most widely used electronic structure technique, d. functional theory (DFT), failing to describe them with std. approxns. Therefore, applications of DFT to systems where dispersion is important have traditionally been of questionable accuracy. However, the last decade has seen a surge of enthusiasm in the DFT community to tackle this problem and in so-doing to extend the applicability of DFT-based methods. Here we discuss, classify, and evaluate some of the promising schemes to emerge in recent years. A brief perspective on the outstanding issues that remain to be resolved and some directions for future research are also provided. (c) 2012 American Institute of Physics.
- 225Ehrlich, S.; Moellmann, J.; Grimme, S. Dispersion-Corrected Density Functional Theory for Aromatic Interactions in Complex Systems Acc. Chem. Res. 2013, 46 (4) 916– 926 DOI: 10.1021/ar3000844Google Scholar225Dispersion-Corrected Density Functional Theory for Aromatic Interactions in Complex SystemsEhrlich, Stephan; Moellmann, Jonas; Grimme, StefanAccounts of Chemical Research (2013), 46 (4), 916-926CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Arom. interactions play a key role in many chem. and biol. systems. However, even if very simple models are chosen, the systems of interest are often too large to be handled with std. wave function theory (WFT). Although d. functional theory (DFT) can easily treat systems of more than 200 atoms, std. semilocal (hybrid) d. functional approxns. fail to describe the London dispersion energy, a factor that is essential for accurate predictions of inter- and intramol. noncovalent interactions. Therefore dispersion-cor. DFT provides a unique tool for the investigation and anal. of a wide range of complex arom. systems. In this Account, we start with an anal. of the noncovalent interactions in simple model dimers of hexafluorobenzene (HFB) and benzene, with a focus on electrostatic and dispersion interactions. The min. for the parallel-displaced dimers of HFB/HFB and HFB/benzene can only be explained when taking into account all contributions to the interaction energy and not by electrostatics alone. By comparison of satd. and arom. model complexes, we show that increased dispersion coeffs. for sp2-hybridized carbon atoms play a major role in arom. stacking. Modern dispersion-cor. DFT yields accurate results (about 5-10% error for the dimerization energy) for the relatively large porphyrin and coronene dimers, systems for which WFT can provide accurate ref. data only with huge computational effort. In this example, it is also demonstrated that new nonlocal, d.-dependent dispersion corrections and atom pairwise schemes mutually agree with each other. The dispersion energy is also important for the complex inter- and intramol. interactions that arise in the mol. crystals of arom. mols. In studies of hexahelicene, dispersion-cor. DFT yields "the right answer for the right reason". By comparison, std. DFT calcns. reproduce intramol. distances quite accurately in single-mol. calcns. while inter- and intramol. distances become too large when dispersion-uncorrected solid-state calcns. are carried out. Dispersion-cor. DFT can fix this problem, and these results are in excellent agreement with exptl. structure and energetic (sublimation) data. Uncorrected treatments do not even yield a bound crystal state. Finally, we present calcns. for the formation of a cationic, quadruply charged dimer of a porphyrin deriv., a case where dispersion is required in order to overcome strong electrostatic repulsion. A combination of dispersion-cor. DFT with an adequate continuum solvation model can accurately reproduce exptl. free assocn. enthalpies in soln. As in the previous examples, consideration of the electrostatic interactions alone does not provide a qual. or quant. correct picture of the interactions of this complex.
- 226DiStasio, R. A., Jr.; Gobre, V. V.; Tkatchenko, A. Many-Body van Der Waals Interactions in Molecules and Condensed Matter J. Phys.: Condens. Matter 2014, 26 (21) 213202 DOI: 10.1088/0953-8984/26/21/213202Google Scholar226Many-body van der Waals interactions in molecules and condensed matterDiStasio, Robert A.; Gobre, Vivekanand V.; Tkatchenko, AlexandreJournal of Physics: Condensed Matter (2014), 26 (21), 213202/1-213202/16, 16 pp.CODEN: JCOMEL; ISSN:0953-8984. (IOP Publishing Ltd.)A review. This work reviews the increasing evidence that many-body van der Waals (vdW) or dispersion interactions play a crucial role in the structure, stability and function of a wide variety of systems in biol., chem. and physics. Starting with the exact expression for the electron correlation energy provided by the adiabatic connection fluctuation-dissipation theorem, we derive both pairwise and many-body interat. methods for computing the long-range dispersion energy by considering a model system of coupled quantum harmonic oscillators within the RPA. By coupling this approach to d. functional theory, the resulting many-body dispersion (MBD) method provides an accurate and efficient scheme for computing the frequency-dependent polarizability and many-body vdW energy in mols. and materials with a finite electronic gap. A select collection of applications are presented that ascertain the fundamental importance of these non-bonded interactions across the spectrum of intermol. (the S22 and S66 benchmark databases), intramol. (conformational energies of alanine tetrapeptide) and supramol. (binding energy of the 'buckyball catcher') complexes, as well as mol. crystals (cohesive energies in oligoacenes). These applications demonstrate that electrodynamic response screening and beyond-pairwise many-body vdW interactions - both captured at the MBD level of theory - play a quant., and sometimes even qual., role in describing the properties considered herein. This work is then concluded with an in-depth discussion of the challenges that remain in the future development of reliable (accurate and efficient) methods for treating many-body vdW interactions in complex materials and provides a road-map for navigating many of the research avenues that are yet to be explored.
- 227Casimir, H.; Polder, D. The Influence of Retardation on the London-van Der Waals Forces Phys. Rev. 1948, 73 (4) 360 DOI: 10.1103/PhysRev.73.360Google Scholar227The influence of retardation on the London-van der Waals forcesCasimir, H. B. G.; Polder, D.Physical Review (1948), 73 (), 360-72CODEN: PHRVAO; ISSN:0031-899X.cf. C.A. 41, 1899i. Math.theoretical. The influence of retardation on the energy of interaction between two neutral atoms is investigated by means of quantum electrodynamics. In the interactions between a neutral atom and a perfectly conducting plane, and between two atoms, it is found that the influence of radiation is described by a monotonically decreasing correction factor which is equal to unity for small distances, R, compared with the wave lengths corresponding to the at. frequencies, and proportional to 1/R when R is large compared with these wave lengths.
- 228Starkschall, G.; Gordon, R. G. Calculation of Coefficients in the Power Series Expansion of the Long-Range Dispersion Force between Atoms J. Chem. Phys. 1972, 56 (6) 2801– 2806 DOI: 10.1063/1.1677610Google ScholarThere is no corresponding record for this reference.
- 229Tang, K.; Toennies, J. P. An Improved Simple Model for the van Der Waals Potential Based on Universal Damping Functions for the Dispersion Coefficients J. Chem. Phys. 1984, 80 (8) 3726– 3741 DOI: 10.1063/1.447150Google Scholar229An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficientsTang, K. T.; Toennies, J. PeterJournal of Chemical Physics (1984), 80 (8), 3726-41CODEN: JCPSA6; ISSN:0021-9606.By using the earlier model (K.T.T. and J.P.T., 1977), a simple expression was derived for the radial-dependent damping functions for the individual dispersion coeffs. C2n for arbitrary even orders 2n. The damping functions are only a function of the Born-Mayer range parameter b, and thus can be applied to all systems for which this is known or can be estd. For H(1S)-H(1S), the results agree with the ab-initio damping functions of A. Koide, et al., (1981). Comparisons with less accurate previous calcns. for other systems also show agreement. By adding a Born-Mayer repulsive term [A exp(-bR)] to the damped dispersion potential, a simple universal expression was obtained for the well region of the atom-atom van der Waals potential with only 5 essential parameters A, b, C6, C8, and C10. The model was tested for the systems: H23Σ, He2, Ar2, NaK3Σ, and LiHg, for which either very precise theor. or exptl. data are available. For each system, the ab-initio dispersion coeffs. together with the parameters ε and Rm were used to det. A and b from the model potential. With these values, the reduced potentials were calcd., and found to agree with the exptl. potentials to better than 1%, and always less than the exptl. uncertainties.
- 230Thakkar, A. J.; Hettema, H.; Wormer, P. E. Abinitio Dispersion Coefficients for Interactions Involving Rare-Gas Atoms J. Chem. Phys. 1992, 97 (5) 3252– 3257 DOI: 10.1063/1.463012Google ScholarThere is no corresponding record for this reference.
- 231Lein, M.; Dobson, J. F.; Gross, E. K. Toward the Description of van Der Waals Interactions within Density Functional Theory J. Comput. Chem. 1999, 20 (1) 12– 22 DOI: 10.1002/(SICI)1096-987X(19990115)20:1<12::AID-JCC4>3.0.CO;2-UGoogle ScholarThere is no corresponding record for this reference.
- 232Kamal, C.; Ghanty, T.; Banerjee, A.; Chakrabarti, A. The van Der Waals Coefficients between Carbon Nanostructures and Small Molecules: A Time-Dependent Density Functional Theory Study J. Chem. Phys. 2009, 131 (16) 164708 DOI: 10.1063/1.3256238Google ScholarThere is no corresponding record for this reference.
- 233Johnson, E. R.; Becke, A. D. A Post-Hartree–Fock Model of Intermolecular Interactions J. Chem. Phys. 2005, 123 (2) 024101 DOI: 10.1063/1.1949201Google Scholar233A post-Hartree-Fock model of intermolecular interactionsJohnson, Erin R.; Becke, Axel D.Journal of Chemical Physics (2005), 123 (2), 024101/1-024101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Intermol. interactions are of great importance in chem. but are difficult to model accurately with computational methods. In particular, Hartree-Fock and std. d.-functional approxns. do not include the physics necessary to properly describe dispersion. These methods are sometimes cor. to account for dispersion by adding a pairwise C6/R6 term, with C6 dispersion coeffs. dependent on the atoms involved. We present a post-Hartree-Fock model in which C6 coeffs. are generated by the instantaneous dipole moment of the exchange hole. This model relies on occupied orbitals only, and involves only one, universal, empirical parameter to limit the dispersion energy at small interat. sepns. The model is extensively tested on isotropic C6 coeffs. of 178 intermol. pairs. It is also applied to the calcn. of the geometries and binding energies of 20 intermol. complexes involving dispersion, dipole-induced dipole, dipole-dipole, and hydrogen-bonding interactions, with remarkably good results.
- 234Becke, A. D.; Johnson, E. R. Exchange-Hole Dipole Moment and the Dispersion Interaction J. Chem. Phys. 2005, 122 (15) 154104 DOI: 10.1063/1.1884601Google Scholar234Exchange-hole dipole moment and the dispersion interactionBecke, Axel D.; Johnson, Erin R.Journal of Chemical Physics (2005), 122 (15), 154104/1-154104/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple model is presented in which the instantaneous dipole moment of the exchange hole is used to generate a dispersion interaction between nonoverlapping systems. The model is easy to implement, requiring no electron correlation (in the usual sense) or time dependence, and has been tested on various at. and mol. pairs. The resulting C6 dispersion coeffs. are remarkably accurate.
- 235Johnson, E. R. Dependence of Dispersion Coefficients on Atomic Environment J. Chem. Phys. 2011, 135 (23) 234109 DOI: 10.1063/1.3670015Google ScholarThere is no corresponding record for this reference.
- 236Misquitta, A. J. Intermolecular Interactions. In Handbook of Computational Chemistry; Springer, 2012; pp 157– 193.Google ScholarThere is no corresponding record for this reference.
- 237Wen, S.; Nanda, K.; Huang, Y.; Beran, G. J. Practical Quantum Mechanics-Based Fragment Methods for Predicting Molecular Crystal Properties Phys. Chem. Chem. Phys. 2012, 14 (21) 7578– 7590 DOI: 10.1039/c2cp23949cGoogle Scholar237Practical quantum mechanics-based fragment methods for predicting molecular crystal propertiesWen, Shuhao; Nanda, Kaushik; Huang, Yuanhang; Beran, Gregory J. O.Physical Chemistry Chemical Physics (2012), 14 (21), 7578-7590CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A review. Significant advances in fragment-based electronic structure methods have created a real alternative to force-field and d. functional techniques in condensed-phase problems such as mol. crystals. This perspective article highlights some of the important challenges in modeling mol. crystals and discusses techniques for addressing them. First, we survey recent developments in fragment-based methods for mol. crystals. Second, we use examples from our own recent research on a fragment-based QM/MM method, the hybrid many-body interaction (HMBI) model, to analyze the phys. requirements for a practical and effective mol. crystal model chem. We demonstrate that it is possible to predict mol. crystal lattice energies to within a couple kJ mol-1 and lattice parameters to within a few percent in small-mol. crystals. Fragment methods provide a systematically improvable approach to making predictions in the condensed phase, which is crit. to making robust predictions regarding the subtle energy differences found in mol. crystals.
- 238Wu, Q.; Yang, W. Empirical Correction to Density Functional Theory for van Der Waals Interactions J. Chem. Phys. 2002, 116 (2) 515– 524 DOI: 10.1063/1.1424928Google Scholar238Empirical correction to density functional theory for van der Waals interactionsWu, Qin; Yang, WeitaoJournal of Chemical Physics (2002), 116 (2), 515-524CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)An empirical method has been designed to account for the van der Waals interactions in practical mol. calcns. with d. functional theory. For each atom pair sepd. at a distance R, the method adds to the d. functional electronic structure calcns. an addnl. attraction energy EvdW = -fd(R)C6R-6, where fd(R) is the damping function which equals to one at large value of R and zero at small value of R. The coeffs. C6 for pair interactions between hydrogen, carbon, nitrogen, and oxygen atoms have been developed in this work by a least-square fitting to the mol. C6 coeffs. obtained from the dipole oscillator strength distribution method by Meath and co-workers. Two forms of the damping functions have been studied, with one dropping to zero at short distances much faster than the other. Four d. functionals have been examd.: Becke's three parameter hybrid functional with the Lee-Yang-Parr correlation functional, Becke's 1988 exchange functional with the LYP correlation functional, Becke's 1988 exchange functional with Perdew and Wang's 1991 (PW91) correlation functional, and PW91 exchange and correlation functional. The method has been applied to three systems where the van der Waals attractions are known to be important: rare-gas diat. mols., stacking of base pairs, and polyalanines' conformation stabilities. The results show that this empirical method, with the damping function dropping to zero smoothly, provides a significant correction to both of the Becke's hybrid functional and the PW91 exchange and correlation functional. Results are comparable with the corresponding second-order Moller-Plesset calcns. in many cases.
- 239Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections Phys. Chem. Chem. Phys. 2008, 10 (44) 6615– 6620 DOI: 10.1039/b810189bGoogle Scholar239Long-range corrected hybrid density functionals with damped atom-atom dispersion correctionsChai, Jeng-Da; Head-Gordon, MartinPhysical Chemistry Chemical Physics (2008), 10 (44), 6615-6620CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We report re-optimization of a recently proposed long-range cor. (LC) hybrid d. functional [J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106] to include empirical atom-atom dispersion corrections. The resulting functional, ωB97X-D yields satisfactory accuracy for thermochem., kinetics, and non-covalent interactions. Tests show that for non-covalent systems, ωB97X-D shows slight improvement over other empirical dispersion-cor. d. functionals, while for covalent systems and kinetics it performs noticeably better. Relative to our previous functionals, such as ωB97X, the new functional is significantly superior for non-bonded interactions, and very similar in performance for bonded interactions.
- 240Liu, Y.; Goddard, W. A. I. A Universal Damping Function for Empirical Dispersion Correction on Density Functional Theory Mater. Trans. 2009, 50 (7) 1664– 1670 DOI: 10.2320/matertrans.MF200911Google Scholar240A universal damping function for empirical dispersion correction on density functional theoryLiu, Yi; Goddard, William A., IIIMaterials Transactions (2009), 50 (7), 1664-1670CODEN: MTARCE; ISSN:1345-9678. (Japan Institute of Metals)A damped London dispersion interaction is generally adopted in empirical dispersion corrections on d. functional theory (DFT), where dispersion parameters are detd. empirically to reproduce correct dispersive interactions after assuming a damping function. The key to a successful dispersion correction is choosing an appropriate damping function. In this work we propose a single universal damping function that can represent several damping functions used in literatures with a few adjustable parameters. This universal damping function provides a unified formula that allows convenient comparison and flexible optimization in dispersion cor. DFT methods. Using the optimized universal damping functions and dispersion parameters, we develop dispersion correction methods for HF, B3LYP and PBE theories. We calc. the dispersive energies accurately for rare gas diat. mols. and benzene dimers with an averaged error <4.1%.
- 241Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu J. Chem. Phys. 2010, 132 (15) 154104 DOI: 10.1063/1.3382344Google Scholar241A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-PuGrimme, Stefan; Antony, Jens; Ehrlich, Stephan; Krieg, HelgeJournal of Chemical Physics (2010), 132 (15), 154104/1-154104/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The method of dispersion correction as an add-on to std. Kohn-Sham d. functional theory (DFT-D) has been refined regarding higher accuracy, broader range of applicability, and less empiricism. The main new ingredients are atom-pairwise specific dispersion coeffs. and cutoff radii that are both computed from first principles. The coeffs. for new eighth-order dispersion terms are computed using established recursion relations. System (geometry) dependent information is used for the first time in a DFT-D type approach by employing the new concept of fractional coordination nos. (CN). They are used to interpolate between dispersion coeffs. of atoms in different chem. environments. The method only requires adjustment of two global parameters for each d. functional, is asymptotically exact for a gas of weakly interacting neutral atoms, and easily allows the computation of at. forces. Three-body nonadditivity terms are considered. The method has been assessed on std. benchmark sets for inter- and intramol. noncovalent interactions with a particular emphasis on a consistent description of light and heavy element systems. The mean abs. deviations for the S22 benchmark set of noncovalent interactions for 11 std. d. functionals decrease by 15%-40% compared to the previous (already accurate) DFT-D version. Spectacular improvements are found for a tripeptide-folding model and all tested metallic systems. The rectification of the long-range behavior and the use of more accurate C6 coeffs. also lead to a much better description of large (infinite) systems as shown for graphene sheets and the adsorption of benzene on an Ag(111) surface. For graphene it is found that the inclusion of three-body terms substantially (by about 10%) weakens the interlayer binding. We propose the revised DFT-D method as a general tool for the computation of the dispersion energy in mols. and solids of any kind with DFT and related (low-cost) electronic structure methods for large systems. (c) 2010 American Institute of Physics.
- 242Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory J. Comput. Chem. 2011, 32 (7) 1456– 1465 DOI: 10.1002/jcc.21759Google Scholar242Effect of the damping function in dispersion corrected density functional theoryGrimme, Stefan; Ehrlich, Stephan; Goerigk, LarsJournal of Computational Chemistry (2011), 32 (7), 1456-1465CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)It is shown by an extensive benchmark on mol. energy data that the math. form of the damping function in DFT-D methods has only a minor impact on the quality of the results. For 12 different functionals, a std. "zero-damping" formula and rational damping to finite values for small interat. distances according to Becke and Johnson (BJ-damping) has been tested. The same (DFT-D3) scheme for the computation of the dispersion coeffs. is used. The BJ-damping requires one fit parameter more for each functional (three instead of two) but has the advantage of avoiding repulsive interat. forces at shorter distances. With BJ-damping better results for nonbonded distances and more clear effects of intramol. dispersion in four representative mol. structures are found. For the noncovalently-bonded structures in the S22 set, both schemes lead to very similar intermol. distances. For noncovalent interaction energies BJ-damping performs slightly better but both variants can be recommended in general. The exception to this is Hartree-Fock that can be recommended only in the BJ-variant and which is then close to the accuracy of cor. GGAs for non-covalent interactions. According to the thermodn. benchmarks BJ-damping is more accurate esp. for medium-range electron correlation problems and only small and practically insignificant double-counting effects are obsd. It seems to provide a phys. correct short-range behavior of correlation/dispersion even with unmodified std. functionals. In any case, the differences between the two methods are much smaller than the overall dispersion effect and often also smaller than the influence of the underlying d. functional. © 2011 Wiley Periodicals, Inc.; J. Comput. Chem., 2011.
- 243Johnson, E. R.; Becke, A. D. A Post-Hartree-Fock Model of Intermolecular Interactions: Inclusion of Higher-Order Corrections J. Chem. Phys. 2006, 124 (17) 174104 DOI: 10.1063/1.2190220Google Scholar243A post-Hartree-Fock model of intermolecular interactions: Inclusion of higher-order correctionsJohnson, Erin R.; Becke, Axel D.Journal of Chemical Physics (2006), 124 (17), 174104/1-174104/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We have previously demonstrated that the dipole moment of the exchange hole can be used to derive intermol. C6 dispersion coeffs. [J. Chem. Phys. 122, 154104 (2005)]. This was subsequently the basis for a novel post-Hartree-Fock model of intermol. interactions [J. Chem. Phys. 123, 024101 (2005)]. In the present work, the model is extended to include higher-order dispersion coeffs. C8 and C10. The extended model performs very well for prediction of intermonomer sepns. and binding energies of 45 van der Waals complexes. In particular, it performs twice as well as basis-set extrapolated MP2 theory for dispersion-bound complexes, with minimal computational cost.
- 244Grimme, S. Accurate Description of van Der Waals Complexes by Density Functional Theory Including Empirical Corrections J. Comput. Chem. 2004, 25 (12) 1463– 1473 DOI: 10.1002/jcc.20078Google Scholar244Accurate description of van der Waals complexes by density functional theory including empirical correctionsGrimme, StefanJournal of Computational Chemistry (2004), 25 (12), 1463-1473CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)An empirical method to account for van der Waals interactions in practical calcns. in the framework of the d. functional theory (termed DFT-D) was tested for a wide variety of mol. complexes. As in previous schemes, the dispersive energy was described by damped interat. potentials of the form C6R-6. The use of pure, gradient-cor. d. functionals (BLYP and PBE), together with the resoln.-of-the-identity (RI) approxn. for the Coulomb operator, allows very efficient computations for large systems. In contrast to the previous work, extended AO basis sets of polarized TZV or QZV quality were employed, which reduced the basis set superposition error to a negligible extend. By using a global scaling factor for the at. C6 coeffs., the functional dependence of the results could be strongly reduced. The "double counting" of correlation effects for strongly bound complexes was found to be insignificant if steep damping functions were employed. The method was applied to a total of 29 complexes of atoms and small mols. (Ne, CH4, NH3, H2O, CH3F, N2, F2, formic acid, ethene, and ethine) with each other and with benzene, to benzene, naphthalene, pyrene, and coronene dimers, the naphthalene trimer, coronene·H2O and four H-bonded and stacked DNA base pairs (AT and GC). In almost all cases, very good agreement with reliable theor. or exptl. results for binding energies and intermol. distances is obtained. For stacked arom. systems and the important base pairs, the DFT-D-BLYP model seems to be even superior to std. MP2 treatments that systematically over-bind. The good results obtained suggest the approach as a practical tool to describe the properties of many important van der Waals systems in chem. Furthermore, the DFT-D data may either be used to calibrate much simpler (e.g., force-field) potentials or the optimized structures can be used as input for more accurate ab initio calcns. of the interaction energies.
- 245Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction J. Comput. Chem. 2006, 27 (15) 1787– 1799 DOI: 10.1002/jcc.20495Google Scholar245Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 246Jurečka, P.; Černỳ, J.; Hobza, P.; Salahub, D. R. Density Functional Theory Augmented with an Empirical Dispersion Term. Interaction Energies and Geometries of 80 Noncovalent Complexes Compared with Ab Initio Quantum Mechanics Calculations J. Comput. Chem. 2007, 28 (2) 555– 569 DOI: 10.1002/jcc.20570Google ScholarThere is no corresponding record for this reference.
- 247Tkatchenko, A.; Scheffler, M. Accurate Molecular van Der Waals Interactions from Ground-State Electron Density and Free-Atom Reference Data Phys. Rev. Lett. 2009, 102 (7) 073005 DOI: 10.1103/PhysRevLett.102.073005Google Scholar247Accurate Molecular Van Der Waals Interactions from Ground-State Electron Density and Free-Atom Reference DataTkatchenko, Alexandre; Scheffler, MatthiasPhysical Review Letters (2009), 102 (7), 073005/1-073005/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We present a parameter-free method for an accurate detn. of long-range van der Waals interactions from mean-field electronic structure calcns. Our method relies on the summation of interat. C6 coeffs., derived from the electron d. of a mol. or solid and accurate ref. data for the free atoms. The mean abs. error in the C6 coeffs. is 5.5% when compared to accurate exptl. values for 1225 intermol. pairs, irresp. of the employed exchange-correlation functional. We show that the effective at. C6 coeffs. depend strongly on the bonding environment of an atom in a mol. Finally, we analyze the van der Waals radii and the damping function in the C6R-6 correction method for d.-functional theory calcns.
- 248Steinmann, S. N.; Corminboeuf, C. A System-Dependent Density-Based Dispersion Correction J. Chem. Theory Comput. 2010, 6 (7) 1990– 2001 DOI: 10.1021/ct1001494Google Scholar248A System-Dependent Density-Based Dispersion CorrectionSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Theory and Computation (2010), 6 (7), 1990-2001CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)D. functional approxns. fail to provide a consistent description of weak mol. interactions arising from small electron d. overlaps. A simple remedy to correct for the missing interactions is to add a posteriori an attractive energy term summed over all atom pairs in the system. The d.-dependent energy correction, presented herein, is applicable to all elements of the periodic table and is easily combined with any electronic structure method, which lacks the accurate treatment of weak interactions. Dispersion coeffs. are computed according to Becke and Johnson's exchange-hole dipole moment (XDM) formalism, thereby depending on the chem. environment of an atom (d., oxidn. state). The long-range ∼R-6 potential is supplemented with higher-order correction terms (∼R-8 and ∼R-10) through the universal damping function of Tang and Toennies. A genuine damping factor depending on (iterative) Hirshfeld (overlap) populations, at. ionization energies, and two adjustable parameters specifically fitted to a given DFT functional is also introduced. The proposed correction, dDXDM, dramatically improves the performance of popular d. functionals. The anal. of 30 (dispersion cor.) d. functionals on 145 systems reveals that dDXDM largely reduces the errors of the parent functionals for both inter- and intramol. interactions. With mean abs. deviations (MADs) of 0.74-0.84 kcal mol-1, PBE-dDXDM, PBE0-dDXDM, and B3LYP-dDXDM outperform the computationally more demanding and most recent functionals such as M06-2X and B2PLYP-D (MAD of 1.93 and 1.06 kcal mol-1, resp.).
- 249Steinmann, S. N.; Corminboeuf, C. Comprehensive Benchmarking of a Density-Dependent Dispersion Correction J. Chem. Theory Comput. 2011, 7 (11) 3567– 3577 DOI: 10.1021/ct200602xGoogle Scholar249Comprehensive Benchmarking of a Density-Dependent Dispersion CorrectionSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Theory and Computation (2011), 7 (11), 3567-3577CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Std. d. functional approxns. cannot accurately describe interactions between nonoverlapping densities. A simple remedy consists in correcting for the missing interactions a posteriori, adding an attractive energy term summed over all atom pairs. The d.-dependent energy correction, dDsC, presented herein, is constructed from dispersion coeffs. computed on the basis of a generalized gradient approxn. to Becke and Johnson's exchange-hole dipole moment formalism. DDsC also relies on an extended Tang and Toennies damping function accounting for charge-overlap effects. The comprehensive benchmarking on 341 diverse reaction energies divided into 18 illustrative test sets validates the robust performance and general accuracy of dDsC for describing various intra- and intermol. interactions. With a total MAD of 1.3 kcal mol-1, B97-dDsC slightly improves the results of M06-2X and B2PLYP-D3 (MAD = 1.4 kcal mol-1 for both) at a lower computational cost. The d. dependence of both the dispersion coeffs. and the damping function makes the approach esp. valuable for modeling redox reactions and charged species in general.
- 250Steinmann, S. N.; Corminboeuf, C. A Generalized-Gradient Approximation Exchange Hole Model for Dispersion Coefficients J. Chem. Phys. 2011, 134 (4) 044117 DOI: 10.1063/1.3545985Google Scholar250A generalized-gradient approximation exchange hole model for dispersion coefficientsSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Physics (2011), 134 (4), 044117/1-044117/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple method for computing accurate d.-dependent dispersion coeffs. is presented. The dispersion coeffs. are modeled by a generalized gradient-type approxn. to Becke and Johnson's exchange hole dipole moment formalism. Our most cost-effective variant, based on a disjoint description of atoms in a mol., gives mean abs. errors in the C6 coeffs. for 90 complexes below 10%. The inclusion of the missing long-range van der Waals interactions in d. functionals using the derived coeffs. in a pair wise correction leads to highly accurate typical noncovalent interaction energies. (c) 2011 American Institute of Physics.
- 251Liu, Y.; Goddard, W. A., III First-Principles-Based Dispersion Augmented Density Functional Theory: From Molecules to Crystals J. Phys. Chem. Lett. 2010, 1 (17) 2550– 2555 DOI: 10.1021/jz100615gGoogle Scholar251First-Principles-Based Dispersion Augmented Density Functional Theory: From Molecules to CrystalsLiu, Yi; Goddard, William A., IIIJournal of Physical Chemistry Letters (2010), 1 (17), 2550-2555CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Std. implementations of d. functional theory (DFT) describe well strongly bound mols. and solids but fail to describe long-range van der Waals attractions. We propose here first-principles-based augmentation to DFT that leads to the proper long-range 1/R6 attraction of the London dispersion while leading to low gradients (small forces) at normal valence distances so that it preserves the accurate geometries and thermochem. of std. DFT methods. The DFT-low gradient (DFT-lg) formula differs from previous DFT-D methods by using a purely attractive dispersion correction while not affecting valence bond distances. We demonstrate here that the DFT-lg model leads to good descriptions for graphite, benzene, naphthalene, and anthracene crystals, using just three parameters fitted to reproduce the full potential curves of high-level ab initio quantum mechanics [CCSD(T)] on gas-phase benzene dimers. The addnl. computational costs for this DFT-lg formalism are negligible.
- 252Kim, H.; Choi, J.-M.; Goddard, W. A., III Universal Correction of Density Functional Theory to Include London Dispersion (up to Lr, Element 103) J. Phys. Chem. Lett. 2012, 3 (3) 360– 363 DOI: 10.1021/jz2016395Google Scholar252Universal Correction of Density Functional Theory to Include London Dispersion (up to Lr, Element 103)Kim, Hyungjun; Choi, Jeong-Mo; Goddard, William A., IIIJournal of Physical Chemistry Letters (2012), 3 (3), 360-363CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Conventional d. functional theory (DFT) fails to describe accurately the London dispersion essential for describing mol. interactions in soft matter (biol. systems, polymers, nucleic acids) and mol. crystals. This has led to several methods in which atom-dependent potentials are added into the Kohn-Sham DFT energy. Some of these corrections were fitted to accurate quantum mech. results, but it will be tedious to det. the appropriate parameters to describe all of the atoms of the periodic table. We propose an alternative approach in which a single parameter in the low-gradient (lg) functional form is combined with the rule-based UFF (universal force-field) nonbond parameters developed for the entire periodic table (up to Lr, Z = 103), named as a DFT-ulg method. We show that DFT-ulg method leads to a very accurate description of the properties for mol. complexes and mol. crystals, providing the means for predicting more accurate weak interactions across the periodic table.
- 253Román-Pérez, G.; Soler, J. M. Efficient Implementation of a van Der Waals Density Functional: Application to Double-Wall Carbon Nanotubes Phys. Rev. Lett. 2009, 103 (9) 096102 DOI: 10.1103/PhysRevLett.103.096102Google Scholar253Efficient Implementation of a van der Waals Density Functional: Application to Double-Wall Carbon NanotubesRoman-Perez, Guillermo; Soler, Jose M.Physical Review Letters (2009), 103 (9), 096102/1-096102/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We present an efficient implementation of the van der Waals d. functional of Dion et al., which expresses the nonlocal correlation energy as a double spatial integral. We factorize the integration kernel and use fast Fourier transforms to evaluate the self-consistent potential, total energy, and at. forces, in O(NlogN) operations. The resulting overhead, for medium and large systems, is a small fraction of the total computational cost, representing a dramatic speedup over the O(N2) evaluation of the double integral. This opens the realm of first-principles simulations to the large systems of interest in soft matter and biomol. problems. We apply the method to calc. the binding energies and the barriers for relative translation and rotation in double-wall carbon nanotubes.
- 254Andersson, Y.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Interactions in Density-Functional Theory Phys. Rev. Lett. 1996, 76 (1) 102– 105 DOI: 10.1103/PhysRevLett.76.102Google ScholarThere is no corresponding record for this reference.
- 255Dobson, J. F.; Dinte, B. P. Constraint Satisfaction in Local and Gradient Susceptibility Approximations: Application to a van Der Waals Density Functional Phys. Rev. Lett. 1996, 76 (11) 1780 DOI: 10.1103/PhysRevLett.76.1780Google ScholarThere is no corresponding record for this reference.
- 256Sato, T.; Tsuneda, T.; Hirao, K. Van Der Waals Interactions Studied by Density Functional Theory Mol. Phys. 2005, 103 (6–8) 1151– 1164 DOI: 10.1080/00268970412331333474Google ScholarThere is no corresponding record for this reference.
- 257Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Density Functional for General Geometries Phys. Rev. Lett. 2004, 92 (24) 246401 DOI: 10.1103/PhysRevLett.92.246401Google Scholar257Van der Waals Density Functional for General GeometriesDion, M.; Rydberg, H.; Schroeder, E.; Langreth, D. C.; Lundqvist, B. I.Physical Review Letters (2004), 92 (24), 246401/1-246401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)A scheme within d. functional theory is proposed that provides a practical way to generalize to unrestricted geometries the method applied with some success to layered geometries [H. Rydberg et al., Phys. Rev. Lett. 91, 126402 (2003)]. It includes van der Waals forces in a seamless fashion. By expansion to second order in a carefully chosen quantity contained in the long-range part of the correlation functional, the nonlocal correlations are expressed in terms of a d.-d. interaction formula. It contains a relatively simple parametrized kernel, with parameters detd. by the local d. and its gradient. The proposed functional is applied to rare gas and benzene dimers, where it is shown to give a realistic description.
- 258Pernal, K.; Podeszwa, R.; Patkowski, K.; Szalewicz, K. Dispersionless Density Functional Theory Phys. Rev. Lett. 2009, 103 (26) 263201 DOI: 10.1103/PhysRevLett.103.263201Google ScholarThere is no corresponding record for this reference.
- 259Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van Der Waals Density Functional J. Phys.: Condens. Matter 2010, 22 (2) 022201 DOI: 10.1088/0953-8984/22/2/022201Google Scholar259Chemical accuracy for the van der Waals density functionalKlimes, Jiri; Bowler, David R.; Michaelides, AngelosJournal of Physics: Condensed Matter (2010), 22 (2), 022201/1-022201/5CODEN: JCOMEL; ISSN:0953-8984. (Institute of Physics Publishing)The non-local van der Waals d. functional (vdW-DF) of Dion et al is a very promising scheme for the efficient treatment of dispersion bonded systems. We show here that the accuracy of vdW-DF can be dramatically improved both for dispersion and hydrogen bonded complexes through the judicious selection of its underlying exchange functional. New and published exchange functionals are identified that deliver much better than chem. accuracy from vdW-DF for the S22 benchmark set of weakly interacting dimers and for water clusters. Improved performance for the adsorption of water on salt is also obtained.
- 260Lee, K.; Murray, É. D.; Kong, L.; Lundqvist, B. I.; Langreth, D. C. Higher-Accuracy van Der Waals Density Functional Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 82 (8) 081101 DOI: 10.1103/PhysRevB.82.081101Google Scholar260Higher-accuracy van der Waals density functionalLee, Kyuho; Murray, Eamonn D.; Kong, Lingzhu; Lundqvist, Bengt I.; Langreth, David C.Physical Review B: Condensed Matter and Materials Physics (2010), 82 (8), 081101/1-081101/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)We propose a second version of the van der Waals d. functional of Dion et al. [Phys. Rev. Lett. 92, 246401 (2004)], employing a more accurate semilocal exchange functional and the use of a large-N asymptote gradient correction in detg. the vdW kernel. The predicted binding energy, equil. sepn., and potential-energy curve shape are close to those of accurate quantum chem. calcns. on 22 duplexes. We anticipate the enabling of chem. accurate calcns. in sparse materials of importance for condensed matter, surface, chem., and biol. physics.
- 261Hamada, I. Van Der Waals Density Functional Made Accurate Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 89 (12) 121103 DOI: 10.1103/PhysRevB.89.121103Google Scholar261van der Waals density functional made accurateHamada, IkutaroPhysical Review B: Condensed Matter and Materials Physics (2014), 89 (12), 121103/1-121103/5CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)I propose a van der Waals d. functional (vdW-DF) that improves upon the description of energetics and geometries of mols., solids, and adsorption systems over the original vdW-DF. The functional is based on the nonlocal correlation for the second version of the vdW-DF and an exchange functional that recovers the second-order gradient expansion approxn. in the slowly varying limit, while reproducing the large d. gradient behavior proposed by Becke. A systematic assessment of the proposed functional is presented, which demonstrates the applicability of the proposed vdW-DF to a wide range of systems.
- 262Vydrov, O. A.; Van Voorhis, T. Nonlocal van Der Waals Density Functional Made Simple Phys. Rev. Lett. 2009, 103 (6) 063004 DOI: 10.1103/PhysRevLett.103.063004Google Scholar262Nonlocal van der Waals Density Functional Made SimpleVydrov, Oleg A.; Van Voorhis, TroyPhysical Review Letters (2009), 103 (6), 063004/1-063004/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We derive a nonlocal correlation functional that adequately describes van der Waals interactions not only in the asymptotic long-range regime but also at short range. Unlike its precursor, developed by Langreth, Lundqvist, and co-workers, the new functional has a simple analytic form, finite for all interelectron sepns., well behaved in the slowly varying d. limit, and generalized to spin-polarized systems.
- 263Vydrov, O. A.; Van Voorhis, T. Improving the Accuracy of the Nonlocal van Der Waals Density Functional with Minimal Empiricism J. Chem. Phys. 2009, 130 (10) 104105 DOI: 10.1063/1.3079684Google ScholarThere is no corresponding record for this reference.
- 264Vydrov, O. A.; Van Voorhis, T. Implementation and Assessment of a Simple Nonlocal van Der Waals Density Functional J. Chem. Phys. 2010, 132 (16) 164113 DOI: 10.1063/1.3398840Google ScholarThere is no corresponding record for this reference.
- 265Vydrov, O. A.; Van Voorhis, T. Nonlocal van Der Waals Density Functional: The Simpler the Better J. Chem. Phys. 2010, 133 (24) 244103 DOI: 10.1063/1.3521275Google Scholar265Nonlocal van der Waals density functional: The simpler the betterVydrov, Oleg A.; Van Voorhis, TroyJournal of Chemical Physics (2010), 133 (24), 244103/1-244103/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We devise a nonlocal correlation energy functional that describes the entire range of dispersion interactions in a seamless fashion using only the electron d. as input. The new functional is considerably simpler than its predecessors of a similar type. The functional has a tractable and robust analytic form that lends itself to efficient self-consistent implementation. When paired with an appropriate exchange functional, our nonlocal correlation model yields accurate interaction energies of weakly-bound complexes, not only near the energy min. but also far from equil. Our model exhibits an outstanding precision at predicting equil. intermonomer sepns. in van der Waals complexes. It also gives accurate covalent bond lengths and atomization energies. Hence the functional proposed in this work is a computationally inexpensive electronic structure tool of broad applicability. (c) 2010 American Institute of Physics.
- 266Sabatini, R.; Gorni, T.; de Gironcoli, S. Nonlocal van Der Waals Density Functional Made Simple and Efficient Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87 (4) 041108 DOI: 10.1103/PhysRevB.87.041108Google Scholar266Nonlocal van der Waals density functional made simple and efficientSabatini, Riccardo; Gorni, Tommaso; de Gironcoli, StefanoPhysical Review B: Condensed Matter and Materials Physics (2013), 87 (4), 041108/1-041108/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)We present a simple revision of the VV10 nonlocal d. functional by Vydrov and Van Voorhis for dispersion interactions. Unlike the original functional our modification allows nonlocal correlation energy and its derivs. to be efficiently evaluated in a plane wave framework along the lines pioneered by Roman-Perez and Soler. Our revised functional maintains the outstanding precision of the original VV10 in noncovalently bound complexes and performs well in representative covalent, ionic, and metallic solids.
- 267Hujo, W.; Grimme, S. Performance of the van Der Waals Density Functional VV10 and (hybrid) GGA Variants for Thermochemistry and Noncovalent Interactions J. Chem. Theory Comput. 2011, 7 (12) 3866– 3871 DOI: 10.1021/ct200644wGoogle Scholar267Performance of the van der Waals Density Functional VV10 and (hybrid)GGA Variants for Thermochemistry and Noncovalent InteractionsHujo, Waldemar; Grimme, StefanJournal of Chemical Theory and Computation (2011), 7 (12), 3866-3871CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal van der Waals d. functional VV10 (Vydrov, O. A.; Van Voorhis, T. J. Chem. Phys.2010, 133, 244103) is tested for the thermochem. properties of 1200 + atoms and mols. in the GMTKN30 database in order to assess its global accuracy. Five GGA and hybrid functionals in unmodified form are augmented by the nonlocal (NL) part of the VV10 functional (one parameter adjusted). The addn. of the NL dispersion energy definitely improves the results of all tested functionals. On the basis of little empiricism and basic phys. insight, DFT-NL can be recommended as a fully electronic, robust electronic structure method.
- 268Aragó, J.; Ortí, E.; Sancho-García, J. C. Nonlocal van Der Waals Approach Merged with Double-Hybrid Density Functionals: Toward the Accurate Treatment of Noncovalent Interactions J. Chem. Theory Comput. 2013, 9 (8) 3437– 3443 DOI: 10.1021/ct4003527Google Scholar268Nonlocal van der Waals Approach Merged with Double-Hybrid Density Functionals: Toward the Accurate Treatment of Noncovalent InteractionsArago, Juan; Orti, Enrique; Sancho-Garcia, Juan C.Journal of Chemical Theory and Computation (2013), 9 (8), 3437-3443CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Noncovalent interactions drive the self-assembly of weakly interacting mol. systems to form supramol. aggregates, which play a major role in nanotechnol. and biochem. In this work, we present a thorough assessment of the performance of different double-hybrid d. functionals (PBE0-DH-NL, revPBE0-DH-NL, B2PLYP-NL, and TPSS0-DH-NL), as well as their parent hybrid and (meta)GGA functionals, in combination with the most modern version of the nonlocal (NL) van der Waals correction. It is shown that this nonlocal correction can be successfully coupled with double-hybrid d. functionals thanks to the short-range attenuation parameter b, which has been optimized against ref. interaction energies of benchmarking mol. complexes (S22 and S66 databases). Among all the double-hybrid functionals evaluated, revPBE0-DH-NL and B2PLYP-NL behave remarkably accurate with mean unsigned errors (MUE) as small as 0.20 kcal/mol for the training sets and in the 0.25-0.42 kcal/mol range for an independent database (NCCE31). They can be thus seen as appropriate functionals to use in a broad no. of applications where noncovalent interactions play an important role. Overall, the nonlocal van der Waals approach combined with last-generation d. functionals is confirmed as an accurate and affordable computational tool for the modeling of weakly bonded mol. systems.
- 269Cooper, V. R. Van Der Waals Density Functional: An Appropriate Exchange Functional Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81 (16) 161104 DOI: 10.1103/PhysRevB.81.161104Google Scholar269Van der Waals density functional: An appropriate exchange functionalCooper, Valentino R.Physical Review B: Condensed Matter and Materials Physics (2010), 81 (16), 161104/1-161104/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)In this Rapid Communication, an exchange functional which is compatible with the nonlocal Rutgers-Chalmers correlation functional [van der Waals d. functional (vdW-DF)] is presented. This functional, when employed with vdW-DF, demonstrates remarkable improvements on intermol. sepn. distances while further improving the accuracy of vdW-DF interaction energies. The key to the success of this three-parameter functional is its redn. in short-range exchange repulsion through matching to the gradient expansion approxn. in the slowly varying/high-d. limit while recovering the large reduced gradient, s, limit set in the revised Perdew-Burke-Ernzerhof (revPBE) exchange functional. This augmented exchange functional could be a soln. to long-standing issues of vdW-DF lending to further applicability of d.-functional theory to the study of relatively large, dispersion bound (van der Waals) complexes.
- 270Vydrov, O. A.; Van Voorhis, T. Benchmark Assessment of the Accuracy of Several van Der Waals Density Functionals J. Chem. Theory Comput. 2012, 8 (6) 1929– 1934 DOI: 10.1021/ct300081yGoogle Scholar270Benchmark Assessment of the Accuracy of Several van der Waals Density FunctionalsVydrov, Oleg A.; Van Voorhis, TroyJournal of Chemical Theory and Computation (2012), 8 (6), 1929-1934CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal correlation functional VV10, developed recently in our group, describes the whole range of dispersion interactions in a seamless and general fashion using only the electron d. as input. The VV10 functional has a simple analytic form that can be adjusted for pairing with the exchange functional of choice. In this paper, we use several benchmark data sets of weakly interacting mol. complexes to test the accuracy of two VV10 variants, differing in their treatment of the exchange component. For the sake of comparison, several other d. functionals suitable for noncovalent interactions were also tested against the same benchmarks. We find that the "default" version of VV10 with semilocal exchange gives very accurate geometries and binding energies for most van der Waals complexes but systematically overbinds hydrogen-bonded complexes. The alternative variant of VV10 with long-range cor. hybrid exchange performs exceptionally well for all types of weak bonding sampled in this study, including hydrogen bonds.
- 271Hujo, W.; Grimme, S. Performance of Non-Local and Atom-Pairwise Dispersion Corrections to DFT for Structural Parameters of Molecules with Noncovalent Interactions J. Chem. Theory Comput. 2013, 9 (1) 308– 315 DOI: 10.1021/ct300813cGoogle Scholar271Performance of Non-Local and Atom-Pairwise Dispersion Corrections to DFT for Structural Parameters of Molecules with Noncovalent InteractionsHujo, Waldemar; Grimme, StefanJournal of Chemical Theory and Computation (2013), 9 (1), 308-315CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal, electron d. dependent dispersion correction of Vydrov and Van Voorhis (Vydrov, O. A.; Van Voorhis, T. J. Chem. Phys.2010, 133, 244103), termed VV10 or DFT-NL, was implemented for structural optimizations of mols. It was tested in combination with the four (hybrid)GGA d. functionals TPSS, TPSS0, B3LYP, and revPBE38 for inter- and intramol. noncovalent interactions (NCI) and compared to results from atom-pairwise dispersion cor. DFT-D3. The methods were applied to a wide range of different problems, namely the S22 and S66 test sets, large transition metal complexes, water hexamer clusters, hexahelicene, and four other difficult cases of intramol. NCI. Crit. interat. distances were computed remarkably accurately by both dispersion corrections compared to theor. or exptl. ref. data and inter- and intramol. interactions were treated on equal footing. The methods can be recommended as reliable and robust tools for geometry optimizations of large systems in which long-range dispersion forces are crucial.
- 272Tran, F.; Hutter, J. Nonlocal van Der Waals Functionals: The Case of Rare-Gas Dimers and Solids J. Chem. Phys. 2013, 138 (20) 204103 DOI: 10.1063/1.4807332Google ScholarThere is no corresponding record for this reference.
- 273Gobre, V. V.; Tkatchenko, A. Scaling Laws for van Der Waals Interactions in Nanostructured Materials Nat. Commun. 2013, 4, 2341 DOI: 10.1038/ncomms3341Google Scholar273Scaling laws for van der Waals interactions in nanostructured materialsGobre Vivekanand V; Tkatchenko AlexandreNature communications (2013), 4 (), 2341 ISSN:.Van der Waals interactions have a fundamental role in biology, physics and chemistry, in particular in the self-assembly and the ensuing function of nanostructured materials. Here we utilize an efficient microscopic method to demonstrate that van der Waals interactions in nanomaterials act at distances greater than typically assumed, and can be characterized by different scaling laws depending on the dimensionality and size of the system. Specifically, we study the behaviour of van der Waals interactions in single-layer and multilayer graphene, fullerenes of varying size, single-wall carbon nanotubes and graphene nanoribbons. As a function of nanostructure size, the van der Waals coefficients follow unusual trends for all of the considered systems, and deviate significantly from the conventionally employed pairwise-additive picture. We propose that the peculiar van der Waals interactions in nanostructured materials could be exploited to control their self-assembly.
- 274Ambrosetti, A.; Alfè, D.; DiStasio, R. A., Jr.; Tkatchenko, A. Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular Systems J. Phys. Chem. Lett. 2014, 5 (5) 849– 855 DOI: 10.1021/jz402663kGoogle Scholar274Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular SystemsAmbrosetti, Alberto; Alfe, Dario; DiStasio, Robert A.; Tkatchenko, AlexandreJournal of Physical Chemistry Letters (2014), 5 (5), 849-855CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Noncovalent interactions are ubiquitous in mol. and condensed-phase environments, and hence a reliable theor. description of these fundamental interactions could pave the way toward a more complete understanding of the microscopic underpinnings for a diverse set of systems in chem. and biol. In this work, we demonstrate that recent algorithmic advances coupled to the availability of large-scale computational resources make the stochastic quantum Monte Carlo approach to solving the Schrodinger equation an optimal contender for attaining "chem. accuracy" (1 kcal/mol) in the binding energies of supramol. complexes of chem. relevance. To illustrate this point, we considered a select set of seven host-guest complexes, representing the spectrum of noncovalent interactions, including dispersion or van der Waals forces, π-π stacking, hydrogen bonding, hydrophobic interactions, and electrostatic (ion-dipole) attraction. A detailed anal. of the interaction energies reveals that a complete theor. description necessitates treatment of terms well beyond the std. London and Axilrod-Teller contributions to the van der Waals dispersion energy.
- 275Šimová, L.; Řezáč, J.; Hobza, P. Convergence of the Interaction Energies in Noncovalent Complexes in the Coupled-Cluster Methods Up to Full Configuration Interaction J. Chem. Theory Comput. 2013, 9 (8) 3420– 3428 DOI: 10.1021/ct4002762Google Scholar275Convergence of the Interaction Energies in Noncovalent Complexes in the Coupled-Cluster Methods Up to Full Configuration InteractionSimova, Lucia; Rezac, Jan; Hobza, PavelJournal of Chemical Theory and Computation (2013), 9 (8), 3420-3428CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The CCSD(T) method stands out among various coupled-cluster (CC) approxns. as the golden std. in computational chem. and is widely and successfully used in the realm of covalent and noncovalent interactions. The CCSD(T) method provides reliable interaction energies, but their surprising accuracy is believed to arise partially from an error compensation. The convergence of the CC expansion has been investigated up to fully iterative pentuple excitations (CCSDTQP); for the smallest eight electron complexes, the full CI calcns. have also been performed. We conclude that the convergence of interaction energy at noncovalent accuracy (0.01 kcal/mol) for the complexes studied is reached already at CCSDTQ or CCSDT(Q) levels. When even higher accuracy (spectroscopic accuracy of 1 cm-1, or 3 cal/mol) is required, then the noniterative CCSDTQ(P) method could be used.
- 276Řezáč, J.; Šimová, L.; Hobza, P. CCSD[T] Describes Noncovalent Interactions Better than the CCSD(T), CCSD(TQ), and CCSDT Methods J. Chem. Theory Comput. 2013, 9 (1) 364– 369 DOI: 10.1021/ct3008777Google Scholar276CCSD[T] Describes Noncovalent Interactions Better than the CCSD(T), CCSD(TQ), and CCSDT MethodsRezac, Jan; Simova, Lucia; Hobza, PavelJournal of Chemical Theory and Computation (2013), 9 (1), 364-369CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The CCSD(T) method is often called the "gold std." of computational chem., because it is one of the most accurate methods applicable to reasonably large mols. It is particularly useful for the description of noncovalent interactions where the inclusion of triple excitations is necessary for achieving a satisfactory accuracy. While it is widely used as a benchmark, the accuracy of CCSD(T) interaction energies has not been reliably quantified yet against more accurate calcns. In this work, we compare the CCSD[T], CCSD(T), and CCSD(TQ) noniterative methods with full CCSDTQ and CCSDT(Q) calcns. We investigate various types of noncovalent complexes [hydrogen-bonded (water dimer, ammonia dimer, water ··· ammonia), dispersion-bound (methane dimer, methane ··· ammonia), and π-π stacked (ethene dimer)] using various coupled-clusters schemes up to CCSDTQ in 6-31G*(0.25), 6-31G**(0.25, 0.15), and aug-cc-pVDZ basis sets. We show that CCSDT(Q) reproduces the CCSDTQ results almost exactly and can thus serve as a benchmark in the cases where CCSDTQ calcns. are not feasible. Surprisingly, the CCSD[T] method provides better agreement with the benchmark values than the other noniterative analogs, CCSD(T) and CCSD(TQ), and even than the much more expensive iterative CCSDT scheme. The CCSD[T] interaction energies differ from the benchmark data by less than 5 cal/mol on av. (for all complexes and all basis sets), whereas the error of CCSD(T) is 9 cal/mol. In larger systems, the difference between these two methods can grow by as much as 0.15 kcal/mol. While this effect can be explained only as an error compensation, the CCSD[T] method certainly deserves more attention in accurate calcns. of noncovalent interactions.
- 277Kong, L.; Bischoff, F. A.; Valeev, E. F. Explicitly Correlated R12/F12 Methods for Electronic Structure Chem. Rev. 2012, 112 (1) 75– 107 DOI: 10.1021/cr200204rGoogle Scholar277Explicitly Correlated R12/F12 Methods for Electronic StructureKong, Liguo; Bischoff, Florian A.; Valeev, Edward F.Chemical Reviews (Washington, DC, United States) (2012), 112 (1), 75-107CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The following topics are discussed: Electron correlation in atoms and mols.(CI wave functions and their symmetries); Explicitly correlated methods and wave functions; Core technol. of modern R12 methods; Coupled-Cluster R12 methods; Multireference R12 methods (R12-MRCI, MRMP2-F12, [2]R12, CASPT2-F12 and MRCI-F12, G-CASSCF, MR CABS Single correction, comparison of MR-R12 Methods).
- 278Burns, L. A.; Marshall, M. S.; Sherrill, C. D. Appointing Silver and Bronze Standards for Noncovalent Interactions: A Comparison of Spin-Component-Scaled (SCS), Explicitly Correlated (F12), and Specialized Wavefunction Approaches J. Chem. Phys. 2014, 141 (23) 234111 DOI: 10.1063/1.4903765Google Scholar278Appointing silver and bronze standards for noncovalent interactions: A comparison of spin-component-scaled (SCS), explicitly correlated (F12), and specialized wavefunction approachesBurns, Lori A.; Marshall, Michael S.; Sherrill, C. DavidJournal of Chemical Physics (2014), 141 (23), 234111/1-234111/21CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A systematic examn. of noncovalent interactions as modeled by wavefunction theory is presented in comparison to gold-std. quality benchmarks available for 345 interaction energies of 49 bimol. complexes. Quantum chem. techniques examd. include spin-component-scaling (SCS) variations on second-order perturbation theory (MP2) [SCS, SCS(N), SCS(MI)] and coupled cluster singles and doubles (CCSD) [SCS, SCS(MI)]; also, method combinations designed to improve dispersion contacts [DW-MP2, MP2C, MP2.5, DW-CCSD(T)-F12]; where available, explicitly correlated (F12) counterparts are also considered. Dunning basis sets augmented by diffuse functions are employed for all accessible ζ-levels; truncations of the diffuse space are also considered. After examn. of both accuracy and performance for 394 model chemistries, SCS(MI)-MP2/cc-pVQZ can be recommended for general use, having good accuracy at low cost and no ill-effects such as imbalance between hydrogen-bonding and dispersion-dominated systems or non-parallelity across dissocn. curves. Moreover, when benchmarking accuracy is desirable but gold-std. computations are unaffordable, this work recommends silver-std. [DW-CCSD(T**)-F12/aug-cc-pVDZ] and bronze-std. [MP2C-F12/aug-cc-pVDZ] model chemistries, which support accuracies of 0.05 and 0.16 kcal/mol and efficiencies of 97.3 and 5.5 h for adenine·thymine, resp. Choice comparisons of wavefunction results with the best symmetry-adapted perturbation theory [T. M. Parker, L. A. Burns, R. M. Parrish, A. G. Ryno, and C. D. Sherrill, J. Chem. Phys.140, 094106 (2014)] and d. functional theory [L. A. Burns, A. Vazquez-Mayagoitia, B. G. Sumpter, and C. D. Sherrill, J. Chem. Phys.134, 084107 (2011)] methods previously studied for these databases are provided for readers' guidance. (c) 2014 American Institute of Physics.
- 279Austin, B. M.; Zubarev, D. Y.; Lester, W. A., Jr. Quantum Monte Carlo and Related Approaches Chem. Rev. 2012, 112 (1) 263– 288 DOI: 10.1021/cr2001564Google Scholar279Quantum Monte Carlo and Related ApproachesAustin, Brian M.; Zubarev, Dmitry Yu.; Lester, William A., Jr.Chemical Reviews (Washington, DC, United States) (2012), 112 (1), 263-288CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The following topics are discussed: Variational Monte Carlo, Fixed-node diffusion Monte Carlo, Self-healing diffusion Monte Carlo, Auxiliary field quantum Monte Carlo, Reptation quantum Monte Carlo, Full CI quantum Monte Carlo, Time-dependent quantum Monte Carlo; Trial electronic wave functions (antisym., backflow transformed, Jastrow), Trial wave function optimization, Effective core potentials; Computational considerations (Linear scaling quantum Monte Carlo, Parallelization and hardware acceleration, Advances in algorithms and software); Applications.
- 280Dubeckỳ, M.; Jurečka, P.; Derian, R.; Hobza, P.; Otyepka, M.; Mitas, L. Quantum Monte Carlo Methods Describe Noncovalent Interactions with Subchemical Accuracy J. Chem. Theory Comput. 2013, 9 (10) 4287– 4292 DOI: 10.1021/ct4006739Google Scholar280Quantum Monte Carlo Methods Describe Noncovalent Interactions with Subchemical AccuracyDubecky, Matus; Jurecka, Petr; Derian, Rene; Hobza, Pavel; Otyepka, Michal; Mitas, LubosJournal of Chemical Theory and Computation (2013), 9 (10), 4287-4292CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An accurate description of noncovalent interaction energies is one of the most challenging tasks in computational chem. To date, nonempirical CCSD-(T)/CBS has been used as a benchmark ref. However, its practical use is limited due to the rapid growth of its computational cost with the system complexity. Here, we show that the fixed-node diffusion Monte Carlo (FN-DMC) method with a more favorable scaling is capable of reaching the CCSD-(T)/CBS within subchem. accuracy (<0.1 kcal/mol) on a testing set of six small noncovalent complexes including the water dimer. In benzene/water, benzene/methane, and the T-shape benzene dimer, FN-DMC provides interaction energies that agree within 0.25 kcal/mol with the best available CCSD-(T)/CBS ests. The demonstrated predictive power of FN-DMC therefore provides new opportunities for studies of the vast and important class of medium/large noncovalent complexes.
- 281Dubeckỳ, M.; Derian, R.; Jurečka, P.; Mitas, L.; Hobza, P.; Otyepka, M. Quantum Monte Carlo for Noncovalent Interactions: An Efficient Protocol Attaining Benchmark Accuracy Phys. Chem. Chem. Phys. 2014, 16 (38) 20915– 20923 DOI: 10.1039/C4CP02093FGoogle ScholarThere is no corresponding record for this reference.
- 282Neese, F.; Wennmohs, F.; Hansen, A. Efficient and Accurate Local Approximations to Coupled-Electron Pair Approaches: An Attempt to Revive the Pair Natural Orbital Method J. Chem. Phys. 2009, 130 (11) 114108 DOI: 10.1063/1.3086717Google Scholar282Efficient and accurate local approximations to coupled-electron pair approaches: An attempt to revive the pair natural orbital methodNeese, Frank; Wennmohs, Frank; Hansen, AndreasJournal of Chemical Physics (2009), 130 (11), 114108/1-114108/18CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Coupled-electron pair approxns. (CEPAs) and coupled-pair functionals (CPFs) have been popular in the 1970s and 1980s and have yielded excellent results for small mols. Recently, interest in CEPA and CPF methods has been renewed. It has been shown that these methods lead to competitive thermochem., kinetic, and structural predictions. They greatly surpass second order Moller-Plesset and popular d. functional theory based approaches in accuracy and are intermediate in quality between CCSD and CCSD(T) in extended benchmark studies. In this work an efficient prodn. level implementation of the closed shell CEPA and CPF methods is reported that can be applied to medium sized mols. in the range of 50-100 atoms and up to about 2000 basis functions. The internal space is spanned by localized internal orbitals. The external space is greatly compressed through the method of pair natural orbitals (PNOs) that was also introduced by the pioneers of the CEPA approaches. Our implementation also makes extended use of d. fitting (or resoln. of the identity) techniques in order to speed up the laborious integral transformations. The method is called local pair natural orbital CEPA (LPNO-CEPA) (LPNO-CPF). The implementation is centered around the concepts of electron pairs and matrix operations. Altogether three cutoff parameters are introduced that control the size of the significant pair list, the av. no. of PNOs per electron pair, and the no. of contributing basis functions per PNO. With the conservatively chosen default values of these thresholds, the method recovers about 99.8% of the canonical correlation energy. This translates to abs. deviations from the canonical result of only a few kcal mol-1. Extended numerical test calcns. demonstrate that LPNO-CEPA (LPNO-CPF) has essentially the same accuracy as parent CEPA (CPF) methods for thermochem., kinetics, weak interactions, and potential energy surfaces but is up to 500 times faster. The method performs best in conjunction with large and flexible basis sets. These results open the way for large-scale chem. applications. (c) 2009 American Institute of Physics.
- 283Neese, F.; Hansen, A.; Liakos, D. G. Efficient and Accurate Approximations to the Local Coupled Cluster Singles Doubles Method Using a Truncated Pair Natural Orbital Basis J. Chem. Phys. 2009, 131 (6) 064103 DOI: 10.1063/1.3173827Google Scholar283Efficient and accurate approximations to the local coupled cluster singles doubles method using a truncated pair natural orbital basisNeese, Frank; Hansen, Andreas; Liakos, Dimitrios G.Journal of Chemical Physics (2009), 131 (6), 064103/1-064103/15CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A prodn. level implementation of the closed-shell local quadratic CI and coupled cluster methods with single and double excitations (QCISD and CCSD) based on the concept of pair natural orbitals (local pair natural orbital LPNO-QCISD and LPNO-CCSD) is reported, evaluated, and discussed. This work is an extension of the earlier developed LPNO coupled-electron pair approxn. (LNPO-CEPA) method and makes extended use of the resoln. of the identity (RI) or d. fitting (DF) approxn. Two variants of each method are compared. The less accurate approxns. (LPNO2-QCISD/LPNO2-CCSD) still recover 98.7%-99.3% of the correlation energy in the given basis and have modest disk space requirements. The more accurate variants (LPNO1-QCISD/LPNO1-CCSD) typically recover 99.75%-99.95% of the correlation energy in the given basis but require the Coulomb and exchange operators with up to two-external indexes to be stored on disk. Both variants have comparable computational efficiency. The convergence of the results with respect to the natural orbital truncation parameter (TCutPNO) has been studied. Extended numerical tests have been performed on abs. and relative correlation energies as function of basis set size and TCutPNO as well as on reaction energies, isomerization energies, and weak intermol. interactions. The results indicate that the errors of the LPNO methods compared to the canonical QCISD and CCSD methods are below 1 kcal/mol with our default thresholds. Finally, some calcns. on larger mols. are reported (ranging from 40-86 atoms) and it is shown that for medium sized mols. the total wall clock time required to complete the LPNO-CCSD calcns. is only two to four times that of the preceding SCF. Thus these methods are highly suitable for large-scale computational chem. applications. Since there are only three thresholds involved that have been given conservative default values, the methods can be confidentially used in a "black-box" fashion in the same way as their canonical counterparts. (c) 2009 American Institute of Physics.
- 284Huntington, L. M.; Nooijen, M. pCCSD: Parameterized Coupled-Cluster Theory with Single and Double Excitations J. Chem. Phys. 2010, 133 (18) 184109 DOI: 10.1063/1.3494113Google ScholarThere is no corresponding record for this reference.
- 285Huntington, L. M. J.; Hansen, A.; Neese, F.; Nooijen, M. Accurate Thermochemistry from a Parameterized Coupled-Cluster Singles and Doubles Model and a Local Pair Natural Orbital Based Implementation for Applications to Larger Systems J. Chem. Phys. 2012, 136 (6) 064101 DOI: 10.1063/1.3682325Google ScholarThere is no corresponding record for this reference.
- 286Hansen, A.; Liakos, D. G.; Neese, F. Efficient and Accurate Local Single Reference Correlation Methods for High-Spin Open-Shell Molecules Using Pair Natural Orbitals J. Chem. Phys. 2011, 135 (21) 214102 DOI: 10.1063/1.3663855Google Scholar286Efficient and accurate local single reference correlation methods for high-spin open-shell molecules using pair natural orbitalsHansen, Andreas; Liakos, Dimitrios G.; Neese, FrankJournal of Chemical Physics (2011), 135 (21), 214102/1-214102/20CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A prodn. level implementation of the high-spin open-shell (spin unrestricted) single ref. coupled pair, quadratic CI and coupled cluster methods with up to doubly excited determinants in the framework of the local pair natural orbital (LPNO) concept is reported. This work is an extension of the closed-shell LPNO methods developed earlier. The internal space is spanned by localized orbitals, while the external space for each electron pair is represented by a truncated PNO expansion. The laborious integral transformation assocd. with the large no. of PNOs becomes feasible through the extensive use of d. fitting (resoln. of the identity (RI)) techniques. Tech. complications arising for the open-shell case and the use of quasi-restricted orbitals for the construction of the ref. determinant are discussed in detail. As in the closed-shell case, only three cutoff parameters control the av. no. of PNOs per electron pair, the size of the significant pair list, and the no. of contributing auxiliary basis functions per PNO. The chosen threshold default values ensure robustness and the results of the parent canonical methods are reproduced to high accuracy. Comprehensive numerical tests on abs. and relative energies as well as timings consistently show that the outstanding performance of the LPNO methods carries over to the open-shell case with minor modifications. Finally, hyperfine couplings calcd. with the variational LPNO-CEPA/1 method, for which a well-defined expectation value type d. exists, indicate the great potential of the LPNO approach for the efficient calcn. of mol. properties. (c) 2011 American Institute of Physics.
- 287Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural Triple Excitations in Local Coupled Cluster Calculations with Pair Natural Orbitals J. Chem. Phys. 2013, 139 (13) 134101 DOI: 10.1063/1.4821834Google Scholar287Natural triple excitations in local coupled cluster calculations with pair natural orbitalsRiplinger, Christoph; Sandhoefer, Barbara; Hansen, Andreas; Neese, FrankJournal of Chemical Physics (2013), 139 (13), 134101/1-134101/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In this work, the extension of the previously developed domain based local pair-natural orbital (DLPNO) based singles- and doubles coupled cluster (DLPNO-CCSD) method to perturbatively include connected triple excitations is reported. The development is based on the concept of triples-natural orbitals that span the joint space of the three pair natural orbital (PNO) spaces of the three electron pairs that are involved in the calcn. of a given triple-excitation contribution. The truncation error is very smooth and can be significantly reduced through extrapolation to the zero threshold. However, the extrapolation procedure does not improve relative energies. The overall computational effort of the method is asymptotically linear with the system size O(N). Actual linear scaling has been confirmed in test calcns. on alkane chains. The accuracy of the DLPNO-CCSD(T) approxn. relative to semicanonical CCSD(T0) is comparable to the previously developed DLPNO-CCSD method relative to canonical CCSD. Relative energies are predicted with an av. error of approx. 0.5 kcal/mol for a challenging test set of medium sized org. mols. The triples correction typically adds 30%-50% to the overall computation time. Thus, very large systems can be treated on the basis of the current implementation. In addn. to the linear C150H302 (452 atoms, >8800 basis functions) we demonstrate the first CCSD(T) level calcn. on an entire protein, Crambin with 644 atoms, and more than 6400 basis functions.
- 288Sparta, M.; Neese, F. Chemical Applications Carried out by Local Pair Natural Orbital Based Coupled-Cluster Methods Chem. Soc. Rev. 2014, 43 (14) 5032– 5041 DOI: 10.1039/c4cs00050aGoogle ScholarThere is no corresponding record for this reference.
- 289Liakos, D. G.; Hansen, A.; Neese, F. Weak Molecular Interactions Studied with Parallel Implementations of the Local Pair Natural Orbital Coupled Pair and Coupled Cluster Methods J. Chem. Theory Comput. 2011, 7 (1) 76– 87 DOI: 10.1021/ct100445sGoogle Scholar289Weak molecular interactions studied with parallel implementations of the local pair natural orbital coupled pair and coupled cluster methodsLiakos, Dimitrios G.; Hansen, Andreas; Neese, FrankJournal of Chemical Theory and Computation (2011), 7 (1), 76-87CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A parallel implementation of the recently developed local pair natural orbital CEPA (LPNO-CEPA/n, n = Version 1, 2, or 3) and the corresponding LPNO coupled cluster method with single- and double excitations (LPNO-CCSD) is described. A detailed anal. alongside pseudocode is presented for the most important computational steps. The scaling with respect to the no. of processors is reasonable and speedups of about 10 with 14 processors have been found in benchmark calcns. (wall-clock time). The most important factor limiting the efficiency of the scaling with respect to the no. of processors is probably the limited bandwidth of the presently prevailing multicore machines. The parallel LPNO methods were applied to study weak intermol. interactions. Initially, the well-established S22 set of mols. was studied. The mean abs. error resulting from the use of the LPNO-CEPA/1 method relative to the most recent CCSD(T) ref. data is found to be 0.24 kcal/mol. Thus, LPNO-CEPA/1 holds great promise for the efficient ab initio treatment of weak intermol. interactions. In order to demonstrate the applicability of the methods to real systems, a two-dimensional potential energy surface for a trimer of 2,4-dihydroxy-3-acetyl-6-Me acetophenone [C11H12O4] (81 atoms, 1296 basis functions, 133 single points) has been calcd. with the LPNO-CEPA/1 method. In this system, a clear distinction can be made between hydrogen bonding and π-π interactions. The global min. on the PES obtained from the calcns. agrees excellently with the exptl. detd. crystal structure. By contrast, popular d. functional methods show no discernible min.
- 290Liakos, D. G.; Neese, F. Improved Correlation Energy Extrapolation Schemes Based on Local Pair Natural Orbital Methods J. Phys. Chem. A 2012, 116 (19) 4801– 4816 DOI: 10.1021/jp302096vGoogle Scholar290Improved Correlation Energy Extrapolation Schemes Based on Local Pair Natural Orbital MethodsLiakos, Dimitrios G.; Neese, FrankJournal of Physical Chemistry A (2012), 116 (19), 4801-4816CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)It is well-known that the basis set limit is difficult to reach in correlated post Hartree-Fock ab initio calcns. One possible route forward is to employ basis set extrapolation schemes. In order to avoid prohibitively expensive calcns., the highest level calcn. (typically based on the "gold std." coupled cluster theory with single, double, and perturbative triple excitations, CCSD(T)) is only performed with the smallest basis set, and the remaining basis set incompleteness is estd. at a lower level of theory, typically second-order Moeller-Plesset perturbation theory (MP2). In this work, we provide a comprehensive investigation of alternative schemes where the MP2 extrapolation is replaced by the coupled-electron pair approxn., version 1 (CEPA/1) or the local pair natural orbital version of this method (LPNO-CEPA/1). It is shown that the MP2 method achieves apparent accuracy only due to error cancellation. Systematically more accurate results at small addnl. computational cost are obtained if the MP2 step is replaced by LPNO-CEPA/1. The errors of LPNO-CEPA/1 relative to canonical CEPA/1 are negligible. Owing to the highly systematic nature of the deviations between canonical and LPNO methods, basis set extrapolation reduces the LPNO errors in the total energies by 1 order of magnitude (∼0.2 kcal/mol) and errors in energy differences to essentially zero. Using the CCSD(T)/LPNO-CEPA/1-based extrapolation scheme, new ref. values are proposed for the recently published S66 set of interaction energies. The deviations between the new values and the original interactions energies are mostly very small but reach values up to 0.3 kcal/mol.
- 291Schwabe, T. Accurate and Fast Treatment of Large Molecular Systems: Assessment of CEPA and pCCSD within the Local Pair Natural Orbital Approximation J. Comput. Chem. 2012, 33 (26) 2067– 2072 DOI: 10.1002/jcc.23042Google ScholarThere is no corresponding record for this reference.
- 292Riplinger, C.; Neese, F. An Efficient and near Linear Scaling Pair Natural Orbital Based Local Coupled Cluster Method J. Chem. Phys. 2013, 138 (3) 034106 DOI: 10.1063/1.4773581Google Scholar292An efficient and near linear scaling pair natural orbital based local coupled cluster methodRiplinger, Christoph; Neese, FrankJournal of Chemical Physics (2013), 138 (3), 034106/1-034106/18CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In previous publications, it was shown that an efficient local coupled cluster method with single- and double excitations can be based on the concept of pair natural orbitals (PNOs) . The resulting local pair natural orbital-coupled-cluster single double (LPNO-CCSD) method has since been proven to be highly reliable and efficient. For large mols., the no. of amplitudes to be detd. is reduced by a factor of 105-106 relative to a canonical CCSD calcn. on the same system with the same basis set. In the original method, the PNOs were expanded in the set of canonical virtual orbitals and single excitations were not truncated. This led to a no. of fifth order scaling steps that eventually rendered the method computationally expensive for large mols. (e.g., >100 atoms). In the present work, these limitations are overcome by a complete redesign of the LPNO-CCSD method. The new method is based on the combination of the concepts of PNOs and projected AOs (PAOs). Thus, each PNO is expanded in a set of PAOs that in turn belong to a given electron pair specific domain. In this way, it is possible to fully exploit locality while maintaining the extremely high compactness of the original LPNO-CCSD wavefunction. No terms are dropped from the CCSD equations and domains are chosen conservatively. The correlation energy loss due to the domains remains below <0.05%, which implies typically 15-20 but occasionally up to 30 atoms per domain on av. The new method has been given the acronym DLPNO-CCSD ("domain based LPNO-CCSD"). The method is nearly linear scaling with respect to system size. The original LPNO-CCSD method had three adjustable truncation thresholds that were chosen conservatively and do not need to be changed for actual applications. In the present treatment, no addnl. truncation parameters have been introduced. Any addnl. truncation is performed on the basis of the three original thresholds. There are no real-space cutoffs. Single excitations are truncated using singles-specific natural orbitals. Pairs are prescreened according to a multipole expansion of a pair correlation energy est. based on local orbital specific virtual orbitals (LOSVs). Like its LPNO-CCSD predecessor, the method is completely of black box character and does not require any user adjustments. It is shown here that DLPNO-CCSD is as accurate as LPNO-CCSD while leading to computational savings exceeding one order of magnitude for larger systems. The largest calcns. reported here featured >8800 basis functions and >450 atoms. In all larger test calcns. done so far, the LPNO-CCSD step took less time than the preceding Hartree-Fock calcn., provided no approxns. have been introduced in the latter. Thus, based on the present development reliable CCSD calcns. on large mols. with unprecedented efficiency and accuracy are realized. (c) 2013 American Institute of Physics.
- 293Sancho-García, J. C.; Aragó, J.; Ortí, E.; Olivier, Y. Obtaining the Lattice Energy of the Anthracene Crystal by Modern yet Affordable First-Principles Methods J. Chem. Phys. 2013, 138 (20) 204304 DOI: 10.1063/1.4806436Google ScholarThere is no corresponding record for this reference.
- 294Grimme, S. Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory Chem. - Eur. J. 2012, 18 (32) 9955– 9964 DOI: 10.1002/chem.201200497Google Scholar294Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional TheoryGrimme, StefanChemistry - A European Journal (2012), 18 (32), 9955-9964, S9955/1-S9955/53CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The equil. assocn. free enthalpies ΔGa for typical supramol. complexes in soln. are calcd. by ab initio quantum chem. methods. Ten neutral and three pos. charged complexes with exptl. ΔGa values in the range 0 to -21 kcal mol-1 (on av. -6 kcal mol-1) are investigated. The theor. approach employs a (non-dynamic) single-structure model, but computes the various energy terms accurately without any special empirical adjustments. Dispersion cor. d. functional theory (DFT-D3) with extended basis sets (triple-ζ and quadruple-ζ quality) is used to det. structures and gas-phase interaction energies (ΔE), the COSMO-RS continuum solvation model (based on DFT data) provides solvation free enthalpies and the remaining ro-vibrational enthalpic/entropic contributions are obtained from harmonic frequency calcns. Low-lying vibrational modes are treated by a free-rotor approxn. The accurate account of London dispersion interactions is mandatory with contributions in the range -5 to -60 kcal mol-1 (up to 200% of ΔE). Inclusion of three-body dispersion effects improves the results considerably. A semi-local (TPSS) and a hybrid d. functional (PW6B95) have been tested. Although the ΔGa values result as a sum of individually large terms with opposite sign (ΔE vs. solvation and entropy change), the approach provides unprecedented accuracy for ΔGa values with errors of only 2 kcal mol-1 on av. Relative affinities for different guests inside the same host are always obtained correctly. The procedure is suggested as a predictive tool in supramol. chem. and can be applied routinely to semirigid systems with 300-400 atoms. The various contributions to binding and enthalpy-entropy compensations are discussed.
- 295Calbo, J.; Ortí, E.; Sancho-García, J. C.; Aragó, J. Accurate Treatment of Large Supramolecular Complexes by Double-Hybrid Density Functionals Coupled with Nonlocal van Der Waals Corrections J. Chem. Theory Comput. 2015, 11 (3) 932– 939 DOI: 10.1021/acs.jctc.5b00002Google Scholar295Accurate Treatment of Large Supramolecular Complexes by Double-Hybrid Density Functionals Coupled with Nonlocal van der Waals CorrectionsCalbo, Joaquin; Orti, Enrique; Sancho-Garcia, Juan C.; Arago, JuanJournal of Chemical Theory and Computation (2015), 11 (3), 932-939CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present a thorough assessment of the performance of some representative double-hybrid d. functionals (revPBE0-DH-NL and B2PLYP-NL) as well as their parent hybrid and GGA counterparts, in combination with the most modern version of the nonlocal (NL) van der Waals correction to describe very large weakly interacting mol. systems dominated by noncovalent interactions. Prior to the assessment, an accurate and homogeneous set of ref. interaction energies was computed for the supramol. complexes constituting the L7 and S12L data sets by using the novel, precise, and efficient DLPNO-CCSD(T) method at the complete basis set limit (CBS). The correction of the basis set superposition error and the inclusion of the deformation energies (for the S12L set) have been crucial for obtaining precise DLPNO-CCSD(T)/CBS interaction energies. Among the d. functionals evaluated, the double-hybrid revPBE0-DH-NL and B2PLYP-NL with the three-body dispersion correction provide remarkably accurate assocn. energies very close to the chem. accuracy. Overall, the NL van der Waals approach combined with proper d. functionals can be seen as an accurate and affordable computational tool for the modeling of large weakly bonded supramol. systems.
- 296Pitoňák, M.; Řezáč, J.; Hobza, P. Spin-Component Scaled Coupled-Clusters Singles and Doubles Optimized towards Calculation of Noncovalent Interactions Phys. Chem. Chem. Phys. 2010, 12 (33) 9611– 9614 DOI: 10.1039/c0cp00158aGoogle ScholarThere is no corresponding record for this reference.
- 297Grimme, S. Improved Second-Order Møller–Plesset Perturbation Theory by Separate Scaling of Parallel-and Antiparallel-Spin Pair Correlation Energies J. Chem. Phys. 2003, 118 (20) 9095– 9102 DOI: 10.1063/1.1569242Google ScholarThere is no corresponding record for this reference.
- 298Szabados, Á. Theoretical Interpretation of Grimme’s Spin-Component-Scaled Second Order Møller-Plesset Theory J. Chem. Phys. 2006, 125 (21) 214105 DOI: 10.1063/1.2404660Google Scholar298Theoretical interpretation of Grimme's spin-component-scaled second order Moller-Plesset theorySzabados, AgnesJournal of Chemical Physics (2006), 125 (21), 214105/1-214105/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)It is shown that spin-component-scaled second order Moller-Plesset theory proposed by Grimme [J. Chem. Phys. 118, 9095 (2003)] can be interpreted as a two-parameter scaling of the zero order Hamiltonian, a generalization of the approach reported by Feenberg [Phys. Rev. 103, 1116 (1956)].
- 299Fink, R. F. Spin-Component-Scaled Møller–Plesset (SCS-MP) Perturbation Theory: A Generalization of the MP Approach with Improved Properties J. Chem. Phys. 2010, 133 (17) 174113 DOI: 10.1063/1.3503041Google ScholarThere is no corresponding record for this reference.
- 300Hill, J. G.; Platts, J. A. Spin-Component Scaling Methods for Weak and Stacking Interactions J. Chem. Theory Comput. 2007, 3 (1) 80– 85 DOI: 10.1021/ct6002737Google Scholar300Spin-Component Scaling Methods for Weak and Stacking InteractionsHill, J. Grant; Platts, James A.Journal of Chemical Theory and Computation (2007), 3 (1), 80-85CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)New scaling parameters are presented for use in the spin-component scaled (SCS) variant of d. fitted local second-order Moller-Plesset perturbation theory (DF-LMP2) that have been optimized for use in evaluating the interaction energy between nucleic acid base pairs. The optimal set of parameters completely neglects the contribution from antiparallel-spin electron pairs to the MP2 energy while scaling the parallel contribution by 1.76. These spin-component scaled for nucleobases (SCSN) parameters are obtained by minimizing, with respect to SCS parameters, the rms interaction energy error relative to the best available literature values, over a set of ten stacked nucleic acid base pairs. The applicability of this scaling to a wide variety of noncovalent interactions is verified through evaluation of a larger set of model complexes, including those dominated by dispersion and electrostatics.
- 301Hill, J. G.; Platts, J. A. Calculating Stacking Interactions in Nucleic Acid Base-Pair Steps Using Spin-Component Scaling and Local Second Order Møller–Plesset Perturbation Theory Phys. Chem. Chem. Phys. 2008, 10 (19) 2785– 2791 DOI: 10.1039/b718691fGoogle ScholarThere is no corresponding record for this reference.
- 302Karton, A.; Tarnopolsky, A.; Lamere, J.-F.; Schatz, G. C.; Martin, J. M. Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical Kinetics J. Phys. Chem. A 2008, 112 (50) 12868– 12886 DOI: 10.1021/jp801805pGoogle Scholar302Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical KineticsKarton, Amir; Tarnopolsky, Alex; Lamere, Jean-Francois; Schatz, George C.; Martin, Jan M. L.Journal of Physical Chemistry A (2008), 112 (50), 12868-12886CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We present a no. of near-exact, nonrelativistic, Born-Oppenheimer ref. data sets for the parametrization of more approx. methods (such as DFT functionals). The data were obtained by means of the W4 ab initio computational thermochem. protocol, which has a 95% confidence interval well below 1 kJ/mol. Our data sets include W4-08, which are total atomization energies of over 100 small mols. that cover varying degrees of non-dynamical correlations, and DBH24-W4, which are W4 theory values for Truhlar's set of 24 representative barrier heights. The usual procedure of comparing calcd. DFT values with exptl. atomization energies is hampered by comparatively large exptl. uncertainties in many exptl. values and compds. errors due to deficiencies in the DFT functional with those resulting from neglect of relativity and finite nuclear mass. Comparison with accurate, explicitly nonrelativistic, ab initio data avoids these issues. We then proceed to explore the performance of B2x-PLYP-type double hybrid functionals for atomization energies and barrier heights. The optimum hybrids for hydrogen-transfer reactions, heavy-atoms transfers, nucleophilic substitutions, and unimol. and recombination reactions are quite different from one another: out of these subsets, the heavy-atom transfer reactions are by far the most sensitive to the percentages of Hartree-Fock-type exchange y and MP2-type correlation x in an (x,y) double hybrid. The (42,72) hybrid B2K-PLYP, as reported in a preliminary communication, represents the best compromise between thermochem. and hydrogen-transfer barriers, while also yielding excellent performance for nucleophilic substitutions. By optimizing for best overall performance on both thermochem. and the DBH24-W4 data set, however, we find a new (36,65) hybrid which we term B2GP-PLYP. At a slight expense in performance for hydrogen-transfer barrier heights and nucleophilic substitutions, we obtain substantially better performance for the other reaction types. Although both B2K-PLYP and B2GP-PLYP are capable of 2 kcal/mol quality thermochem., B2GP-PLYP appears to be the more robust toward non-dynamical correlation and strongly polar character. We addnl. find that double-hybrid functionals display excellent performance for such problems as hydrogen bonding, prototype late transition metal reactions, pericyclic reactions, prototype cumulene-polyacetylene system, and weak interactions.
- 303Distasio, R. A., Jr.; Head-Gordon, M. Optimized Spin-Component Scaled Second-Order Møller-Plesset Perturbation Theory for Intermolecular Interaction Energies Mol. Phys. 2007, 105 (8) 1073– 1083 DOI: 10.1080/00268970701283781Google ScholarThere is no corresponding record for this reference.
- 304Marchetti, O.; Werner, H.-J. Accurate Calculations of Intermolecular Interaction Energies Using Explicitly Correlated Coupled Cluster Wave Functions and a Dispersion-Weighted MP2 Method J. Phys. Chem. A 2009, 113 (43) 11580– 11585 DOI: 10.1021/jp9059467Google ScholarThere is no corresponding record for this reference.
- 305Riley, K. E.; Platts, J. A.; Řezáč, J.; Hobza, P.; Hill, J. G. Assessment of the Performance of MP2 and MP2 Variants for the Treatment of Noncovalent Interactions J. Phys. Chem. A 2012, 116 (16) 4159– 4169 DOI: 10.1021/jp211997bGoogle ScholarThere is no corresponding record for this reference.
- 306Takatani, T.; Hohenstein, E. G.; Sherrill, C. D. Improvement of the Coupled-Cluster Singles and Doubles Method via Scaling Same-and Opposite-Spin Components of the Double Excitation Correlation Energy J. Chem. Phys. 2008, 128 (12) 124111 DOI: 10.1063/1.2883974Google ScholarThere is no corresponding record for this reference.
- 307Jung, Y.; Lochan, R. C.; Dutoi, A. D.; Head-Gordon, M. Scaled Opposite-Spin Second Order Møller–Plesset Correlation Energy: An Economical Electronic Structure Method J. Chem. Phys. 2004, 121 (20) 9793– 9802 DOI: 10.1063/1.1809602Google ScholarThere is no corresponding record for this reference.
- 308Lochan, R. C.; Jung, Y.; Head-Gordon, M. Scaled Opposite Spin Second Order Møller-Plesset Theory with Improved Physical Description of Long-Range Dispersion Interactions J. Phys. Chem. A 2005, 109 (33) 7598– 7605 DOI: 10.1021/jp0514426Google ScholarThere is no corresponding record for this reference.
- 309Takatani, T.; Sherrill, C. D. Performance of Spin-Component-Scaled Møller–Plesset Theory (SCS-MP2) for Potential Energy Curves of Noncovalent Interactions Phys. Chem. Chem. Phys. 2007, 9 (46) 6106– 6114 DOI: 10.1039/b709669kGoogle ScholarThere is no corresponding record for this reference.
- 310Antony, J.; Grimme, S. Is Spin-Component Scaled Second-Order Møller-Plesset Perturbation Theory an Appropriate Method for the Study of Noncovalent Interactions in Molecules? J. Phys. Chem. A 2007, 111 (22) 4862– 4868 DOI: 10.1021/jp070589pGoogle ScholarThere is no corresponding record for this reference.
- 311Bachorz, R. A.; Bischoff, F. A.; Höfener, S.; Klopper, W.; Ottiger, P.; Leist, R.; Frey, J. A.; Leutwyler, S. Scope and Limitations of the SCS-MP2 Method for Stacking and Hydrogen Bonding Interactions Phys. Chem. Chem. Phys. 2008, 10 (19) 2758– 2766 DOI: 10.1039/b718494hGoogle ScholarThere is no corresponding record for this reference.
- 312King, R. A. On the Accuracy of Spin-Component-Scaled Perturbation Theory (SCS-MP2) for the Potential Energy Surface of the Ethylene Dimer Mol. Phys. 2009, 107 (8–12) 789– 795 DOI: 10.1080/00268970802641242Google ScholarThere is no corresponding record for this reference.
- 313Grabowski, I.; Fabiano, E.; Sala, F. D. A Simple Non-Empirical Procedure for Spin-Component-Scaled MP2 Methods Applied to the Calculation of the Dissociation Energy Curve of Noncovalently-Interacting Systems Phys. Chem. Chem. Phys. 2013, 15 (37) 15485– 15493 DOI: 10.1039/c3cp51431eGoogle ScholarThere is no corresponding record for this reference.
- 314Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van Der Waals Complexes Chem. Rev. 1994, 94 (7) 1887– 1930 DOI: 10.1021/cr00031a008Google Scholar314Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van der Waals ComplexesJeziorski, Bogumil; Moszynski, Robert; Szalewicz, KrzysztofChemical Reviews (Washington, DC, United States) (1994), 94 (7), 1887-930CODEN: CHREAY; ISSN:0009-2665.The topics reviewed with 445 refs. include: polarization theory; exchange effects; multipole expansion of the interaction energy; charge-overlap effects and bipolar expansion of polarization energies; the intramonomer electron correlation problem and many-body formulation of symmetry-adapted perturbation theory; and applications.
- 315Misquitta, A. J.; Podeszwa, R.; Jeziorski, B.; Szalewicz, K. Intermolecular Potentials Based on Symmetry-Adapted Perturbation Theory with Dispersion Energies from Time-Dependent Density-Functional Calculations J. Chem. Phys. 2005, 123 (21) 214103 DOI: 10.1063/1.2135288Google Scholar315Intermolecular potentials based on symmetry-adapted perturbation theory with dispersion energies from time-dependent density-functional calculationsMisquitta, Alston J.; Podeszwa, Rafal; Jeziorski, Bogumil; Szalewicz, KrzysztofJournal of Chemical Physics (2005), 123 (21), 214103/1-214103/14CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Recently, three of us have proposed a method [Phys. Rev. Lett. 91, 33201 (2003)] for an accurate calcn. of the dispersion energy utilizing frequency-dependent d. susceptibilities of monomers obtained from time-dependent d.-functional theory (DFT). In the present paper, we report numerical calcns. for the helium, neon, water, and carbon dioxide dimers and show that for a wide range of intermonomer sepns., including the van der Waals and short-range repulsion regions, the method provides dispersion energies with accuracies comparable to those that can be achieved using the current most sophisticated wave-function methods. If the dispersion energy is combined with (i) the electrostatic and first-order exchange interaction energies as defined in symmetry-adapted perturbation theory (SAPT) but computed using monomer Kohn-Sham (KS) determinants, and (ii) the induction energy computed using the coupled KS static response theory, (iii) the exchange-induction and exchange-dispersion energies computed using KS orbitals and orbital energies, the resulting method, denoted by SAPT(DFT), produces very accurate total interaction potentials. For the helium dimer, the only system with nearly exact benchmark values, SAPT(DFT) reproduces the interaction energy to within about 2% at the min. and to a similar accuracy for all other distances ranging from the strongly repulsive to the asymptotic region. For the remaining systems investigated by us, the quality of the SAPT(DFT) interaction energies is so high that these energies may actually be more accurate than the best available results obtained with wave-function techniques. At the same time, SAPT(DFT) is much more computationally efficient than any method previously used for calcg. the dispersion and other interaction energy components at this level of accuracy.
- 316Hesselmann, A.; Jansen, G.; Schütz, M. Interaction Energy Contributions of H-Bonded and Stacked Structures of the AT and GC DNA Base Pairs from the Combined Density Functional Theory and Intermolecular Perturbation Theory Approach J. Am. Chem. Soc. 2006, 128 (36) 11730– 11731 DOI: 10.1021/ja0633363Google ScholarThere is no corresponding record for this reference.
- 317Rezáč, J.; Riley, K. E.; Hobza, P. S66: A Well-Balanced Database of Benchmark Interaction Energies Relevant to Biomolecular Structures J. Chem. Theory Comput. 2011, 7 (8) 2427– 2438 DOI: 10.1021/ct2002946Google Scholar317S66: A Well-balanced Database of Benchmark Interaction Energies Relevant to Biomolecular StructuresRezac, Jan; Riley, Kevin E.; Hobza, PavelJournal of Chemical Theory and Computation (2011), 7 (8), 2427-2438CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)With numerous new quantum chem. methods being developed in recent years and the promise of even more new methods to be developed in the near future, it is clearly crit. that highly accurate, well-balanced, ref. data for many different at. and mol. properties be available for the parametrization and validation of these methods. One area of research that is of particular importance in many areas of chem., biol., and material science is the study of noncovalent interactions. Because these interactions are often strongly influenced by correlation effects, it is necessary to use computationally expensive high-order wave function methods to describe them accurately. Here, the authors present a large new database of interaction energies calcd. using an accurate CCSD(T)/CBS scheme. Data are presented for 66 mol. complexes, at their ref. equil. geometries and at 8 points systematically exploring their dissocn. curves; in total, the database contains 594 points: 66 at equil. geometries, and 528 in dissocn. curves. The data set is designed to cover the most common types of noncovalent interactions in biomols., while keeping a balanced representation of dispersion and electrostatic contributions. The data set is therefore well suited for testing and development of methods applicable to bioorg. systems. In addn. to the benchmark CCSD(T) results, the authors also provide decompns. of the interaction energies by DFT-SAPT calcns. The data set was used to test several correlated QM methods, including those parametrized specifically for noncovalent interactions. Among these, the SCS-MI-CCSD method outperforms all other tested methods, with a root-mean-square error of 0.08 kcal/mol for the S66 data set.
- 318Sinnokrot, M. O.; Sherrill, C. D. Substituent Effects in Π–π Interactions: Sandwich and T-Shaped Configurations J. Am. Chem. Soc. 2004, 126 (24) 7690– 7697 DOI: 10.1021/ja049434aGoogle Scholar318Substituent Effects in π-π Interactions: Sandwich and T-Shaped ConfigurationsSinnokrot, Mutasem Omar; Sherrill, C. DavidJournal of the American Chemical Society (2004), 126 (24), 7690-7697CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Sandwich and T-shaped configurations of benzene dimer, benzene-phenol, benzene-toluene, benzene-fluorobenzene, and benzene-benzonitrile are studied by coupled-cluster theory to elucidate how substituents tune π-π interactions. All substituted sandwich dimers bind more strongly than benzene dimer, whereas the T-shaped configurations bind more or less favorably depending on the substituent. Symmetry-adapted perturbation theory (SAPT) indicates that electrostatic, dispersion, induction, and exchange-repulsion contributions are all significant to the overall binding energies, and all but induction are important in detg. relative energies. Models of π-π interactions based solely on electrostatics, such as the Hunter-Sanders rules, do not seem capable of explaining the energetic ordering of the dimers considered.
- 319Sure, R.; Grimme, S. Corrected Small Basis Set Hartree-Fock Method for Large Systems J. Comput. Chem. 2013, 34 (19) 1672– 1685 DOI: 10.1002/jcc.23317Google Scholar319Corrected small basis set Hartree-Fock method for large systemsSure, Rebecca; Grimme, StefanJournal of Computational Chemistry (2013), 34 (19), 1672-1685CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A quantum chem. method based on a Hartree-Fock calcn. with a small Gaussian AO basis set is presented. Its main area of application is the computation of structures, vibrational frequencies, and noncovalent interaction energies in huge mol. systems. The method is suggested as a partial replacement of semiempirical approaches or d. functional theory (DFT) in particular when self-interaction errors are acute. In order to get accurate results three phys. plausible atom pair-wise correction terms are applied for London dispersion interactions (D3 scheme), basis set superposition error (gCP scheme), and short-ranged basis set incompleteness effects. In total nine global empirical parameters are used. This so-called Hartee-Fock-3c (HF-3c) method is tested for geometries of small org. mols., interaction energies and geometries of noncovalently bound complexes, for supramol. systems, and protein structures. In the majority of realistic test cases good results approaching large basis set DFT quality are obtained at a tiny fraction of computational cost. © 2013 Wiley Periodicals, Inc.
- 320Goerigk, L.; Collyer, C. A.; Reimers, J. R. Recommending Hartree–Fock Theory with London-Dispersion and Basis-Set-Superposition Corrections for the Optimization or Quantum Refinement of Protein Structures J. Phys. Chem. B 2014, 118 (50) 14612– 14626 DOI: 10.1021/jp510148hGoogle ScholarThere is no corresponding record for this reference.
- 321Brandenburg, J.; Grimme, S. Dispersion Corrected Hartree-Fock and Density Functional Theory for Organic Crystal Structure Prediction Top. Curr. Chem. 2013, 345, 1 DOI: 10.1007/128_2013_488Google ScholarThere is no corresponding record for this reference.
- 322Brandenburg, J. G.; Hochheim, M.; Bredow, T.; Grimme, S. Low-Cost Quantum Chemical Methods for Noncovalent Interactions J. Phys. Chem. Lett. 2014, 5 (24) 4275– 4284 DOI: 10.1021/jz5021313Google Scholar322Low-Cost Quantum Chemical Methods for Noncovalent InteractionsBrandenburg, Jan Gerit; Hochheim, Manuel; Bredow, Thomas; Grimme, StefanJournal of Physical Chemistry Letters (2014), 5 (24), 4275-4284CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)A review. The efficient and reasonably accurate description of noncovalent interactions is important for various areas of chem., ranging from supramol. host-guest complexes and biomol. applications to the challenging task of crystal structure prediction. While London dispersion inclusive d. functional theory (DFT-D) can be applied, faster "low-cost" methods are required for large-scale applications. In this Perspective, we present the state-of-the-art of minimal basis set, semiempirical mol.-orbital-based methods. Various levels of approxns. are discussed based either on canonical Hartree-Fock or on semilocal d. functionals. The performance for intermol. interactions is examd. on various small to large mol. complexes and org. solids covering many different chem. groups and interaction types. We put the accuracy of low-cost methods into perspective by comparing with first-principle d. functional theory results. The mean unsigned deviations of binding energies from ref. data are typically 10-30%, which is only two times larger than those of DFT-D. In particular, for neutral or moderately polar systems, many of the tested methods perform very well, while at the same time, computational savings of up to 2 orders of magnitude can be achieved.
- 323Grimme, S.; Brandenburg, J. G.; Bannwarth, C.; Hansen, A. Consistent Structures and Interactions by Density Functional Theory with Small Atomic Orbital Basis Sets J. Chem. Phys. 2015, 143 (5) 054107 DOI: 10.1063/1.4927476Google Scholar323Consistent structures and interactions by density functional theory with small atomic orbital basis setsGrimme, Stefan; Brandenburg, Jan Gerit; Bannwarth, Christoph; Hansen, AndreasJournal of Chemical Physics (2015), 143 (5), 054107/1-054107/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A d. functional theory (DFT) based composite electronic structure approach is proposed to efficiently compute structures and interaction energies in large chem. systems. It is based on the well-known and numerically robust Perdew-Burke-Ernzerhoff (PBE) generalized-gradient-approxn. in a modified global hybrid functional with a relatively large amt. of non-local Fock-exchange. The orbitals are expanded in Ahlrichs-type valence-double zeta AO Gaussian basis sets, which are available for many elements. In order to correct for the basis set superposition error (BSSE) and to account for the important long-range London dispersion effects, our well-established atom-pairwise potentials are used. In the design of the new method, particular attention has been paid to an accurate description of structural parameters in various covalent and non-covalent bonding situations as well as in periodic systems. Together with the recently proposed three-fold cor. (3c) Hartree-Fock method, the new composite scheme (termed PBEh-3c) represents the next member in a hierarchy of "low-cost" electronic structure approaches. They are mainly free of BSSE and account for most interactions in a phys. sound and asymptotically correct manner. PBEh-3c yields good results for thermochem. properties in the huge GMTKN30 energy database. Furthermore, the method shows excellent performance for non-covalent interaction energies in small and large complexes. For evaluating its performance on equil. structures, a new compilation of std. test sets is suggested. These consist of small (light) mols., partially flexible, medium-sized org. mols., mols. comprising heavy main group elements, larger systems with long bonds, 3d-transition metal systems, non-covalently bound complexes (S22 and S66×8 sets), and peptide conformations. For these sets, overall deviations from accurate ref. data are smaller than for various other tested DFT methods and reach that of triple-zeta AO basis set second-order perturbation theory (MP2/TZ) level at a tiny fraction of computational effort. Periodic calcns. conducted for mol. crystals to test structures (including cell vols.) and sublimation enthalpies indicate very good accuracy competitive to computationally more involved plane-wave based calcns. PBEh-3c can be applied routinely to several hundreds of atoms on a single processor and it is suggested as a robust "high-speed" computational tool in theor. chem. and physics. (c) 2015 American Institute of Physics.
- 324Stewart, J. J. Optimization of Parameters for Semiempirical Methods I. Method J. Comput. Chem. 1989, 10 (2) 209– 220 DOI: 10.1002/jcc.540100208Google Scholar324Optimization of parameters for semiempirical methods. I. MethodStewart, James J. P.Journal of Computational Chemistry (1989), 10 (2), 209-20CODEN: JCCHDD; ISSN:0192-8651.A new method for obtaining optimized parameters for semiempirical methods is developed and applied to the MNDO method. The method uses derivs. of calcd. values for properties with respect to adjustable parameters to obtain the optimized values of parameters. The large increase in speed is a result of using a simple series expression for calcd. values of properties rather than employing full semiempirical calcns. With this optimization procedure, the rate-detg. step for parameterizing elements changes from the mechanics of parameterization to the assembling of exptl. ref. data.
- 325Weber, W.; Thiel, W. Orthogonalization Corrections for Semiempirical Methods Theor. Chem. Acc. 2000, 103 (6) 495– 506 DOI: 10.1007/s002149900083Google Scholar325Orthogonalization corrections for semiempirical methodsWeber, Wolfgang; Thiel, WalterTheoretical Chemistry Accounts (2000), 103 (6), 495-506CODEN: TCACFW; ISSN:1432-881X. (Springer-Verlag)Based on a general discussion of orthogonalization effects, two new one-electron orthogonalization corrections are derived to improve existing semiempirical models at the NDDO level. The first one accounts for valence-shell orthogonalization effects on the resonance integrals, while the second one includes the dominant repulsive core-valence interactions through an effective core potential. The corrections for the resonance integrals consist of three-center terms which incorporate stereodiscriminating properties into the two-center matrix elements of the core Hamiltonian. They provide a better description of conformational properties, which is rationalized qual. and demonstrated through numerical calcns. on small model systems. The proposed corrections are part of a new general-purpose semiempirical method which will be described elsewhere.
- 326Stewart, J. J. Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters J. Mol. Model. 2013, 19 (1) 1– 32 DOI: 10.1007/s00894-012-1667-xGoogle Scholar326Optimization of parameters for semiempirical methods VI: more modifications to the NDDO approximations and re-optimization of parametersStewart, James J. P.Journal of Molecular Modeling (2013), 19 (1), 1-32CODEN: JMMOFK; ISSN:0948-5023. (Springer)Modern semiempirical methods are of sufficient accuracy when used in the modeling of mols. of the same type as used as ref. data in the parameterization. Outside that subset, however, there is an abundance of evidence that these methods are of very limited utility. In an attempt to expand the range of applicability, a new method called PM7 has been developed. PM7 was parameterized using exptl. and high-level ab initio ref. data, augmented by a new type of ref. data intended to better define the structure of parameter space. The resulting method was tested by modeling crystal structures and heats of formation of solids. Two changes were made to the set of approxns.: a modification was made to improve the description of noncovalent interactions, and two minor errors in the NDDO formalism were rectified. Av. unsigned errors (AUEs) in geometry and ΔH f for PM7 were reduced relative to PM6; for simple gas-phase org. systems, the AUE in bond lengths decreased by about 5 % and the AUE in ΔH f decreased by about 10 %; for org. solids, the AUE in ΔH f dropped by 60 % and the redn. was 33.3 % for geometries. A two-step process (PM7-TS) for calcg. the heights of activation barriers has been developed. Using PM7-TS, the AUE in the barrier heights for simple org. reactions was decreased from values of 12.6 kcal/mol-1 in PM6 and 10.8 kcal/mol-1 in PM7 to 3.8 kcal/mol-1. The origins of the errors in NDDO methods have been examd., and were found to be attributable to inadequate and inaccurate ref. data. This conclusion provides insight into how these methods can be improved.
- 327Tuttle, T.; Thiel, W. OM X-D: Semiempirical Methods with Orthogonalization and Dispersion Corrections. Implementation and Biochemical Application Phys. Chem. Chem. Phys. 2008, 10 (16) 2159– 2166 DOI: 10.1039/b718795eGoogle ScholarThere is no corresponding record for this reference.
- 328Thiel, W. Semiempirical Quantum–chemical Methods Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4 (2) 145– 157 DOI: 10.1002/wcms.1161Google Scholar328Semiempirical quantum-chemical methodsThiel, WalterWiley Interdisciplinary Reviews: Computational Molecular Science (2014), 4 (2), 145-157CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)The semiempirical methods of quantum chem. are reviewed, with emphasis on established NDDO-based methods (MNDO, AM1, PM3) and on the more recent orthogonalization-cor. methods (OM1, OM2, OM3). After a brief historical overview, the methodol. is presented in nontech. terms, covering the underlying concepts, parameterization strategies, and computational aspects, as well as linear scaling and hybrid approaches. The application section addresses selected recent benchmarks and surveys ground-state and excited-state studies, including recent OM2-based excited-state dynamics investigations.
- 329Řezáč, J.; Fanfrlík, J.; Salahub, D.; Hobza, P. Semiempirical Quantum Chemical PM6 Method Augmented by Dispersion and H-Bonding Correction Terms Reliably Describes Various Types of Noncovalent Complexes J. Chem. Theory Comput. 2009, 5 (7) 1749– 1760 DOI: 10.1021/ct9000922Google Scholar329Semiempirical Quantum Chemical PM6 Method Augmented by Dispersion and H-Bonding Correction Terms Reliably Describes Various Types of Noncovalent ComplexesRezac, Jan; Fanfrlik, Jindrich; Salahub, Dennis; Hobza, PavelJournal of Chemical Theory and Computation (2009), 5 (7), 1749-1760CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Because of its construction and parametrization for more than 80 elements, the semiempirical quantum chem. PM6 method is superior to other similar methods. Despite its advantages, however, the PM6 method fails for the description of noncovalent interactions, specifically the dispersion energy and H-bonding. Upon inclusion of correction terms for dispersion and H-bonding, the performance of the method was found to be dramatically improved. The former correction included two parameters in the damping function that were parametrized to reproduce the benchmark interaction energies [CCSD(T)/complete basis set (CBS) limit] of the dispersion-bonded complexes from the S22 data set. The latter correction was parametrized on an extended set of H-bonded stabilization energies detd. at the MP2/cc-pVTZ level. The resulting PM6-DH method was tested on the S22 data set, for which chem. accuracy (error < 1 kcal/mol) was achieved, and also on the JSCH2005 set, for which significant improvement over the original PM6 method was also obtained. Implementation of anal. gradients allows very efficient geometry optimization, which, for all complexes, provides better agreement with the benchmark data. Excellent results were also achieved for small peptides, and here again, chem. accuracy was obtained (i.e., the error with respect to CCSD(T)/CBS results was smaller than 1 kcal/mol). The performance of the technique was finally demonstrated on extended complexes, namely, the porphine dimer and various graphene models with DNA bases and base pairs, where the PM6-DH stabilization energies agree very well with available benchmark data obtained with DFT-D, SCS-MP2, and MP2.5 methods. The PM6-DH calcns. are very efficient and can be routinely applied for systems of up to 1000 atoms. For nonarom. systems, the use of a linear scaling version of the SCF procedure based on localized orbitals speeds up the method significantly and allows one to investigate systems with several thousand atoms. The method can thus replace force fields, which face basic problems for the description of quantum effects, in many applications.
- 330Korth, M.; Pitoňák, M.; Řezáč, J.; Hobza, P. A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical Methods J. Chem. Theory Comput. 2010, 6 (1) 344– 352 DOI: 10.1021/ct900541nGoogle Scholar330A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical MethodsKorth, Martin; Pitonak, Michal; Rezac, Jan; Hobza, PavelJournal of Chemical Theory and Computation (2010), 6 (1), 344-352CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Semiempirical methods could offer a feasible compromise between ab initio and empirical approaches for the calcn. of large mols. with biol. relevance. A key problem for attempts in this direction is the rather bad performance of current semiempirical methods for noncovalent interactions, esp. hydrogen-bonding. On the basis of the recently introduced PM6-DH method, which includes empirical corrections for dispersion (D) and hydrogen-bond (H) interactions, we have developed an improved and transferable H-bonding correction for semiempirical quantum chem. methods. The performance of the improved correction is evaluated for PM6, AM1, OM3, and SCC-DFTB (enhanced by std. empirical dispersion corrections) with several test sets for noncovalent interactions and is shown to reach the quality of current DFT-D approaches for these types of problems.
- 331Porezag, D.; Frauenheim, T.; Köhler, T.; Seifert, G.; Kaschner, R. Construction of Tight-Binding-like Potentials on the Basis of Density-Functional Theory: Application to Carbon Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 51 (19) 12947 DOI: 10.1103/PhysRevB.51.12947Google Scholar331Construction of tight-binding-like potentials on the basis of density-functional theory: application to carbonPorezag, D.; Frauenheim, Th.; Koehler, Th.; Seifert, G.; Kaschner, R.Physical Review B: Condensed Matter (1995), 51 (19), 12947-57CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The authors present a d.-functional-based scheme for detg. the necessary parameters of common nonorthogonal tight-binding (TB) models within the framework of the LCAO formalism using the local-d. approxn. (LDA). By only considering two-center integrals the Hamiltonian and overlap matrix elements are calcd. out of suitable input densities and potentials rather than fitted to exptl. data. Anal. functions can be derived for the C-C, C-H, and H-H Hamiltonians and overlap matrix elements. The usual short-range repulsive potential appearing in most TB models is fitted to self-consistent calcns. performed within the LDA. The calcn. of forces is easy and allows an application of the method to mol.-dynamics simulations. Despite its extreme simplicity, the method is transferable to complex carbon and hydrocarbon systems. The detn. of equil. geometries, total energies, and vibrational modes of carbon clusters, hydrocarbon mols., and solid-state modifications of carbon yield results showing an overall good agreement with more sophisticated methods.
- 332Seifert, G.; Porezag, D.; Frauenheim, T. Calculations of Molecules, Clusters, and Solids with a Simplified LCAO-DFT-LDA Scheme Int. J. Quantum Chem. 1996, 58 (2) 185– 192 DOI: 10.1002/(SICI)1097-461X(1996)58:2<185::AID-QUA7>3.3.CO;2-BGoogle ScholarThere is no corresponding record for this reference.
- 333Elstner, M.; Hobza, P.; Frauenheim, T.; Suhai, S.; Kaxiras, E. Hydrogen Bonding and Stacking Interactions of Nucleic Acid Base Pairs: A Density-Functional-Theory Based Treatment J. Chem. Phys. 2001, 114 (12) 5149– 5155 DOI: 10.1063/1.1329889Google Scholar333Hydrogen bonding and stacking interactions of nucleic acid base pairs: A density-functional-theory based treatmentElstner, Marcus; Hobza, Pavel; Frauenheim, Thomas; Suhai, Sandor; Kaxiras, EfthimiosJournal of Chemical Physics (2001), 114 (12), 5149-5155CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We extend an approx. d. functional theory (DFT) method for the description of long-range dispersive interactions which are normally neglected by construction, irresp. of the correlation function applied. An empirical formula, consisting of an R-6 term is introduced, which is appropriately damped for short distances; the corresponding C6 coeff., which is calcd. from exptl. at. polarizabilities, can be consistently added to the total energy expression of the method. We apply this approx. DFT plus dispersion energy method to describe the hydrogen bonding and stacking interactions of nucleic acid base pairs. Comparison to MP2/6-31G*(0.25) results shows that the method is capable of reproducing hydrogen bonding as well as the vertical and twist dependence of the interaction energy very accurately.
- 334Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, T.; Suhai, S.; Seifert, G. Self-Consistent-Charge Density-Functional Tight-Binding Method for Simulations of Complex Materials Properties Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58 (11) 7260 DOI: 10.1103/PhysRevB.58.7260Google Scholar334Self-consistent-charge density-functional tight-binding method for simulations of complex materials propertiesElstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, Th.; Suhai, S.; Seifert, G.Physical Review B: Condensed Matter and Materials Physics (1998), 58 (11), 7260-7268CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)We outline details about an extension of the tight-binding (TB) approach to improve total energies, forces, and transferability. The method is based on a second-order expansion of the Kohn-Sham total energy in d.-functional theory (DFT) with respect to charge-d. fluctuations. The zeroth-order approach is equiv. to a common std. non-self-consistent (TB) scheme, while at second-order a transparent, parameter-free, and readily calculable expression for generalized Hamiltonian matrix elements may be derived. These are modified by a self-consistent redistribution of Mulliken charges (SCC). Besides the usual "band structure" and short-range repulsive terms the final approx. Kohn-Sham energy addnl. includes a Coulomb interaction between charge fluctuations. At large distances this accounts for long-range electrostatic forces between two point charges and approx. includes self-interaction contributions of a given atom if the charges are located at one and the same atom. We apply the new SCC scheme to problems where deficiencies within the non-SCC std. TB approach become obvious. We thus considerably improve transferability.
- 335Kunsági-Máté, S.; Ortmann, E.; Kollár, L.; Szabó, K.; Nikfardjam, M. P. Effect of Ferrous and Ferric Ions on Copigmentation in Model Solutions J. Mol. Struct. 2008, 891 (1–3) 471– 474 DOI: 10.1016/j.molstruc.2008.04.036Google ScholarThere is no corresponding record for this reference.
- 336Bondi, A. J. Van Der Waals Volumes and Radii J. Phys. Chem. 1964, 68 (3) 441– 451 DOI: 10.1021/j100785a001Google Scholar336van der Waals volumes and radiiBondi, A.Journal of Physical Chemistry (1964), 68 (3), 441-51CODEN: JPCHAX; ISSN:0022-3654.Intermol. van der Waals radii of the nonmetallic elements were assembled into a list of recommended values for vol. calcns. These values were arrived at by selecting from the most reliable x-ray diffraction data those which could be reconciled with crystal d. at 0°K. (to give reasonable packing d.), gas kinetic collision cross section, crit. d., and with liquid state properties. A qual. understanding of the nature of van der Waals radii is provided by correlation with the de Broglie wavelength of the outermost valence electron. Tentative values for the van der Waals radii of metallic elements in organometallic compds. are proposed. A list of increments for the vol. of mols. impenetrable to thermal collision, the so-called van der Waals vol., and of the corresponding increments in area per mol. is given.
- 337
As seen in previous sections, we are aware that in principle, CCSD(T), or other methods such as QMC, are more desirable. However, such methods are currently too computationally demanding to be applied to real-world copigmentation systems.
There is no corresponding record for this reference. - 338Freitas, A. A.; Shimizu, K.; Dias, L. G.; Quina, F. H. A Computational Study of Substituted Flavylium Salts and Their Quinonoidal Conjugate-Bases: S0→S1 Electronic Transition, Absolute pKa and Reduction Potential Calculations by DFT and Semiempirical Methods J. Braz. Chem. Soc. 2007, 18 (8) 1537– 1546 DOI: 10.1590/S0103-50532007000800014Google Scholar338A computational study of substituted flavylium salts and their quinonoidal conjugate-bases: S0→S1 electronic transition, absolute pKa and reduction potential calculations by DFT and semiempirical methodsFreitas, Adilson A.; Shimizu, K.; Dias, Luis G.; Quina, Frank H.Journal of the Brazilian Chemical Society (2007), 18 (8), 1537-1546CODEN: JOCSET; ISSN:0103-5053. (Sociedade Brasileira de Quimica)The electronic transitions for flavylium cations and quinonoidal bases of 17 substituted flavylium salts have been studied at semiempirical and DFT (d. functional theory) levels. Solvent effect on electronic spectra was included by Polarizable Continuum Model, PCM. We assigned longest-wavelength absorption maxima to HOMO→LUMO transition. Both levels of theory gave good results for electronic transitions of flavylium cations whereas only TDDFT-PCM calcns. could be used for electronic transitions of their quinonoidal bases. We also performed abs. pKa calcns. of nine flavylium salts at DFT level. The pKa calcd. values by our PCM parameterization gave excellent results with mean abs. deviation less than a half of one pKa unit. One-electron redn. potentials were carried out for 5 flavylium cations at DFT level. The theor. results found were in good agreement with exptl. values after adjustment for a systematic deviation.
- 339Sakata, K.; Saito, N.; Honda, T. Ab Initio Study of Molecular Structures and Excited States in Anthocyanidins Tetrahedron 2006, 62 (15) 3721– 3731 DOI: 10.1016/j.tet.2006.01.081Google Scholar339Ab initio study of molecular structures and excited states in anthocyanidinsSakata, Ken; Saito, Norio; Honda, ToshioTetrahedron (2006), 62 (15), 3721-3731CODEN: TETRAB; ISSN:0040-4020. (Elsevier B.V.)The structural and electronic characters of four types of hydroxyl group-substituted anthocyanidins (pelargonidin, cyanidin, delphinidin, and aurantinidin) were examd. using quantum chem. calcns. For these cationic mols., both the planar and non-planar structures in the electronic ground state were detd. at the B3LYP/D95 level of theory. We revealed that the planar structure is slightly more stable than the non-planar structure for each mol. For the optimized planar structures, single excitation-CI (SE-CI) based on the RHF (RHF) wave function was evaluated and the electronic character in the low-excited states was discussed in terms of the MO theory. Symmetry adapted cluster (SAC)/SAC-CI calcns. were also carried out to est. the excitation energies precisely. The results showed that hydroxylation of the Ph group causes a change in the excitation energies without taking the solvent effects into account. The results are in agreement with spectral expts. and previous MO calcns.
- 340Anouar, E. H.; Gierschner, J.; Duroux, J.-L.; Trouillas, P. UV/Visible Spectra of Natural Polyphenols: A Time-Dependent Density Functional Theory Study Food Chem. 2012, 131 (1) 79– 89 DOI: 10.1016/j.foodchem.2011.08.034Google Scholar340UV/Visible spectra of natural polyphenols: A time-dependent density functional theory studyAnouar, El Hassane; Gierschner, Johannes; Duroux, Jean-Luc; Trouillas, PatrickFood Chemistry (2012), 131 (1), 79-89CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)In addn. to their numerous biol. activities, natural and hemisynthetic polyphenols contribute to the large variety of colors (from red to violet) in nature (e.g., fruit, vegetables, leaves and petals). In order to understand the color variation attributed to the multitude of chem. structures of this wide class of compds., time-dependent d. functional quantum-chem. calcns. at the B3P86/6-311+G(d,p) level of theory appears as a relevant and efficient tool. The UV/Vis properties of 33 polyphenols were systematically investigated, including mainly flavonoids, isoflavonoids and flavonolignans. On the basis of MO anal. we established the structure-property relationship, inter alia showing the role of π orbital (de-)localization, mesomeric (+M) effects of hydroxyl groups and structural modification of the mol. backbone. The results might help in the future, for example, for the prediction of novel hemisynthetic compds.
- 341Millot, M.; Di Meo, F.; Tomasi, S.; Boustie, J.; Trouillas, P. Photoprotective Capacities of Lichen Metabolites: A Joint Theoretical and Experimental Study. J. Photochem. Photobiol., B 111, 17– 26. DOI: 10.1016/j.jphotobiol.2012.03.005Google ScholarThere is no corresponding record for this reference.
- 342Carvalho, A. R. F.; Oliveira, J.; De Freitas, V.; Mateus, N.; Melo, A. Unusual Color Change of Vinylpyranoanthocyanin–Phenolic Pigments J. Agric. Food Chem. 2010, 58 (7) 4292– 4297 DOI: 10.1021/jf904246gGoogle ScholarThere is no corresponding record for this reference.
- 343Quartarolo, A. D.; Russo, N. A Computational Study (TDDFT and RICC2) of the Electronic Spectra of Pyranoanthocyanins in the Gas Phase and Solution J. Chem. Theory Comput. 2011, 7 (4) 1073– 1081 DOI: 10.1021/ct2000974Google Scholar343A Computational Study (TDDFT and RICC2) of the Electronic Spectra of Pyranoanthocyanins in the Gas Phase and SolutionQuartarolo, Angelo Domenico; Russo, NinoJournal of Chemical Theory and Computation (2011), 7 (4), 1073-1081CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The conformational structures and UV-vis absorption electronic spectra of a class of derived anthocyanin mols. (pyranoanthocyanins) have been investigated mainly by means of d. functional (DFT) and time-dependent DFT methods. Pyranoanthocyanins are natural pigments present in aged wines and absorb at shorter wavelengths (around 500 nm) than the parent anthocyanin compds., giving an orange-brown colored soln. The investigated mols. are derived from the reaction of glycosylated malvidin, peonidin, and petunidin with enolizable mols. (acetaldehyde and pyruvic acid) and vinyl derivs. During wine storage, the concn. of pyranoanthocyanins increases with time, and anal. measurements (e.g., UV-vis spectroscopy) can characterize aged wines by color anal. The prediction of absorption electronic spectra from TDDFT results, with the inclusion of water bulk solvation effects through the conductor-like polarizable continuum model, gives an abs. mean deviation from exptl. absorption maxima of 0.1 eV and a good reprodn. of the spectra line shape over the visible range of the spectrum. TDDFT calcd. excitation energies agree with those obtained from ab initio multireference coupled cluster with the resoln. of identity approxn. (RICC2) methods, calcd. at DFT gas-phase geometries.
- 344Milián-Medina, B.; Gierschner, J. Computational Design of Low Singlet–triplet Gap All-Organic Molecules for OLED Application Org. Electron. 2012, 13 (6) 985– 991 DOI: 10.1016/j.orgel.2012.02.010Google Scholar344Computational design of low singlet-triplet gap all-organic molecules for OLED applicationMilian-Medina, Begona; Gierschner, JohannesOrganic Electronics (2012), 13 (6), 985-991CODEN: OERLAU; ISSN:1566-1199. (Elsevier B.V.)It was recently reported that external quantum efficiency in org. LEDs can be substantially enhanced when triplet excitons are harvested through upconversion by E-type delayed fluorescence in materials with small singlet-triplet energy gap ΔEST, based on donor-acceptor (DA) chromophores. Furthermore, org. solar cells (OSCs) might profit from such materials in order to reduce recombination losses. However, targeted design rules for such materials are missing up to now. In this paper, we follow a facile (TD-)DFT-based computational design concept by engineering the fragment frontier orbitals in DA systems. The calcns. show that optimized systems with very small ΔEST in the range of kT can be achieved by balancing the energetic offset between fragment MOs as well as through the nature of the DA connector. Application in OLED will addnl. require small non-radiative rates, which recommends large bandgap materials. Utilization in polymeric DA systems with small ΔEST in OSCs requires the full exploration of the chain length dependence of the resp. oligomers.
- 345Dreuw, A.; Weisman, J. L.; Head-Gordon, M. Long-Range Charge-Transfer Excited States in Time-Dependent Density Functional Theory Require Non-Local Exchange J. Chem. Phys. 2003, 119 (6) 2943– 2946 DOI: 10.1063/1.1590951Google Scholar345Long-range charge-transfer excited states in time-dependent density functional theory require non-local exchangeDreuw, Andreas; Weisman, Jennifer L.; Head-Gordon, MartinJournal of Chemical Physics (2003), 119 (6), 2943-2946CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The electrostatic attraction between the sepd. charges in long-range excited charge-transfer states originates from the non-local Hartree-Fock exchange potential and is, thus, a non-local property. Present-day time-dependent d. functional theory employing local exchange-correlation functionals does not capture this effect and therefore fails to describe charge-transfer excited states correctly. A hybrid method that is qual. correct is described.
- 346Dreuw, A.; Head-Gordon, M. Single-Reference Ab Initio Methods for the Calculation of Excited States of Large Molecules Chem. Rev. 2005, 105 (11) 4009– 4037 DOI: 10.1021/cr0505627Google Scholar346Single-Reference ab Initio Methods for the Calculation of Excited States of Large MoleculesDreuw, Andreas; Head-Gordon, MartinChemical Reviews (Washington, DC, United States) (2005), 105 (11), 4009-4037CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review of single-ref. ab initio excited state methods, which are applicable to large mols. and do not explicitly include correlation through the ground-state wave function. CIS, TDHF, and TDDFT are rigorously introduced by outlining their derivations and theor. footing, where special emphasis is put on their relations to each other. Different methods for the anal. of complicated electronically excited states are reviewed and comparatively discussed. The applicability of the presented methods is outlined, their limitations are show, and out their strengths and weaknesses are pointed out.
- 347Casida, M. E.; Gutierrez, F.; Guan, J.; Gadea, F.-X.; Salahub, D.; Daudey, J.-P. Charge-Transfer Correction for Improved Time-Dependent Local Density Approximation Excited-State Potential Energy Curves: Analysis within the Two-Level Model with Illustration for H2 and LiH J. Chem. Phys. 2000, 113 (17) 7062– 7071 DOI: 10.1063/1.1313558Google Scholar347Charge-transfer correction for improved time-dependent local density approximation excited-state potential energy curves: Analysis within the two-level model with illustration for H2 and LiHCasida, Mark E.; Gutierrez, Fabien; Guan, Jingang; Gadea, Florent-Xavier; Salahub, Dennis; Daudey, Jean-PierreJournal of Chemical Physics (2000), 113 (17), 7062-7071CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Time-dependent d.-functional theory (TDDFT) is an increasingly popular approach for calcg. mol. excitation energies. However, the TDDFT lowest triplet excitation energy, ωT, of a closed-shell mol. often falls rapidly to zero and then becomes imaginary at large internuclear distances. We show that this unphys. behavior occurs because ωT2 must become neg. wherever symmetry breaking lowers the energy of the ground state soln. below that of the symmetry unbroken soln. We use the fact that the ΔSCF method gives a qual. correct first triplet excited state to derive a "charge-transfer correction" (CTC) for the time-dependent local d. approxn. (TDLDA) within the two-level model and the Tamm-Dancoff approxn. (TDA). Although this correction would not be needed for the exact exchange-correlation functional, it is evidently important for a correct description of mol. excited state potential energy surfaces in the TDLDA. As a byproduct of our anal., we show why TDLDA and LDA ΔSCF excitation energies are often very similar near the equil. geometries. The reasoning given here is fairly general and it is expected that similar corrections will be needed in the case of generalized gradient approxns. and hybrid functionals.
- 348Yanai, T.; Tew, D. P.; Handy, N. C. A New Hybrid Exchange–correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP) Chem. Phys. Lett. 2004, 393 (1–3) 51– 57 DOI: 10.1016/j.cplett.2004.06.011Google Scholar348A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP)Yanai, Takeshi; Tew, David P.; Handy, Nicholas C.Chemical Physics Letters (2004), 393 (1-3), 51-57CODEN: CHPLBC; ISSN:0009-2614. (Elsevier Science B.V.)A new hybrid exchange-correlation functional named CAM-B3LYP is proposed. It combines the hybrid qualities of B3LYP and the long-range correction presented by Tawada et al. [J. Chem. Phys., in press]. We demonstrate that CAM-B3LYP yields atomization energies of similar quality to those from B3LYP, while also performing well for charge transfer excitations in a dipeptide model, which B3LYP underestimates enormously. The CAM-B3LYP functional comprises of 0.19 Hartree-Fock (HF) plus 0.81 Becke 1988 (B88) exchange interaction at short-range, and 0.65 HF plus 0.35 B88 at long-range. The intermediate region is smoothly described through the std. error function with parameter 0.33.
- 349Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–atom Dispersion Corrections Phys. Chem. Chem. Phys. 2008, 10 (44) 6615– 6620 DOI: 10.1039/b810189bGoogle Scholar349Long-range corrected hybrid density functionals with damped atom-atom dispersion correctionsChai, Jeng-Da; Head-Gordon, MartinPhysical Chemistry Chemical Physics (2008), 10 (44), 6615-6620CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We report re-optimization of a recently proposed long-range cor. (LC) hybrid d. functional [J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106] to include empirical atom-atom dispersion corrections. The resulting functional, ωB97X-D yields satisfactory accuracy for thermochem., kinetics, and non-covalent interactions. Tests show that for non-covalent systems, ωB97X-D shows slight improvement over other empirical dispersion-cor. d. functionals, while for covalent systems and kinetics it performs noticeably better. Relative to our previous functionals, such as ωB97X, the new functional is significantly superior for non-bonded interactions, and very similar in performance for bonded interactions.
- 350Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Double-Hybrid Density Functionals J. Chem. Phys. 2009, 131 (17) 174105 DOI: 10.1063/1.3244209Google Scholar350Long-range corrected double-hybrid density functionalsChai, Jeng-Da; Head-Gordon, MartinJournal of Chemical Physics (2009), 131 (17), 174105/1-174105/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We extend the range of applicability of our previous long-range cor. (LC) hybrid functional, ωB97X, with a nonlocal description of electron correlation, inspired by second-order Moller-Plesset (many-body) perturbation theory. This LC "double-hybrid" d. functional, denoted as ωB97X-2, is fully optimized both at the complete basis set limit (using 2-point extrapolation from calcns. using triple and quadruple zeta basis sets), and also sep. using the somewhat less expensive 6-311++G(3df,3pd) basis. On independent test calcns. (as well as training set results), ωB97X-2 yields high accuracy for thermochem., kinetics, and noncovalent interactions. In addn., owing to its high fraction of exact Hartree-Fock exchange, ωB97X-2 shows significant improvement for the systems where self-interaction errors are severe, such as sym. homonuclear radical cations. (c) 2009 American Institute of Physics.
- 351Caricato, M.; Trucks, G. W.; Frisch, M. J.; Wiberg, K. B. Oscillator Strength: How Does TDDFT Compare to EOM-CCSD? J. Chem. Theory Comput. 2011, 7 (2) 456– 466 DOI: 10.1021/ct100662nGoogle Scholar351Oscillator Strength: How Does TDDFT Compare to EOM-CCSD?Caricato, Marco; Trucks, Gary W.; Frisch, Michael J.; Wiberg, Kenneth B.Journal of Chemical Theory and Computation (2011), 7 (2), 456-466CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We compare a large variety of d. functionals against the equation of motion coupled cluster singles and doubles (EOM-CCSD) method for the calcn. of oscillator strengths. Valence and Rydberg states are considered for a test set composed of 11 small org. mols. In our previous work, the same systems and methods were tested against exptl. results for the excitation energies. The results from this investigation confirm our previous findings, i.e., that there is a large difference between the functionals. For the oscillator strength, the av. best agreement with EOM-CCSD is provided by CAM-B3LYP followed by LC-ωPBE and, to a lesser extent, B3P86 and LC-BLYP.
- 352Beyhan, S. M.; Götz, A. W.; Ariese, F.; Visscher, L.; Gooijer, C. Computational Study on the Anomalous Fluorescence Behavior of Isoflavones J. Phys. Chem. A 2011, 115 (9) 1493– 1499 DOI: 10.1021/jp109059eGoogle ScholarThere is no corresponding record for this reference.
- 353Malcıoğlu, O. B.; Calzolari, A.; Gebauer, R.; Varsano, D.; Baroni, S. Dielectric and Thermal Effects on the Optical Properties of Natural Dyes: A Case Study on Solvated Cyanin J. Am. Chem. Soc. 2011, 133 (39) 15425– 15433 DOI: 10.1021/ja201733vGoogle ScholarThere is no corresponding record for this reference.
- 354Petrone, A.; Cerezo, J.; Ferrer, F. J. A.; Donati, G.; Improta, R.; Rega, N.; Santoro, F. Absorption and Emission Spectral Shapes of a Prototype Dye in Water by Combining Classical/Dynamical and Quantum/Static Approaches J. Phys. Chem. A 2015, 119 (21) 5426– 5438 DOI: 10.1021/jp510838mGoogle ScholarThere is no corresponding record for this reference.
- 355Sinnecker, S.; Rajendran, A.; Klamt, A.; Diedenhofen, M.; Neese, F. Calculation of Solvent Shifts on Electronic G-Tensors with the Conductor-Like Screening Model (COSMO) and Its Self-Consistent Generalization to Real Solvents (Direct COSMO-RS) J. Phys. Chem. A 2006, 110, 2235– 2245 DOI: 10.1021/jp056016zGoogle Scholar355Calculation of Solvent Shifts on Electronic g-Tensors with the Conductor-Like Screening Model (COSMO) and Its Self-Consistent Generalization to Real Solvents (Direct COSMO-RS)Sinnecker, Sebastian; Rajendran, Arivazhagan; Klamt, Andreas; Diedenhofen, Michael; Neese, FrankJournal of Physical Chemistry A (2006), 110 (6), 2235-2245CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The conductor-like screening model (COSMO) was used to investigate the solvent influence on electronic g-values of org. radicals. The previously studied di-Ph nitric oxide and di-tert-Bu nitric oxide radicals were taken as test cases. The calcns. employed spin-unrestricted d. functional theory and the BP and B3LYP d. functionals. The g-tensors were calcd. as mixed second deriv. properties with respect to the external magnetic field and the electron magnetic moment. The first-order response of the Kohn-Sham orbitals with respect to the external magnetic field was detd. through the coupled-perturbed DFT approach. The spin-orbit coupling operator was treated using an accurate multicenter spin-orbit mean-field (SOMF) approach. Provided that important hydrogen bonds are explicitly modeled by a supermol. approach and that the basis set is sufficiently satd., the COSMO calcns. lead to accurate predictions of isotropic g-shifts with deviations of not more than 100 ppm relative to expt. Very accurate results were obtained by employing a recently developed self-consistent modification of the COSMO method to real solvents (COSMO-RS), which we briefly introduce in this paper as direct COSMO-RS (D-COSMO-RS). This model gives isotropic g-shifts of similar high accuracy for water without using the supermol. approach. This is an important result because it solves many of the problems assocd. with the supermol. approach such as local min. and the choice of a suitable model system. Thus, the self-consistent D-COSMO-RS incorporates some specific solvation effects into continuum models, in particular it appears to successfully model the effects of hydrogen bonding. Although not yet widely validated, this opens a novel approach for the calcn. of properties which so far only could be calcd. by the inclusion of explicit solvent mols. in continuum solvation methods.
- 356Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models Chem. Rev. 2005, 105, 2999– 3093 DOI: 10.1021/cr9904009Google Scholar356Quantum Mechanical Continuum Solvation ModelsTomasi, Jacopo; Mennucci, Benedetta; Cammi, RobertoChemical Reviews (Washington, DC, United States) (2005), 105 (8), 2999-3093CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review.
- 357Ge, X.; Timrov, I.; Binnie, S.; Biancardi, A.; Calzolari, A.; Baroni, S. Accurate and Inexpensive Prediction of the Color Optical Properties of Anthocyanins in Solution J. Phys. Chem. A 2015, 119 (16) 3816– 3822 DOI: 10.1021/acs.jpca.5b01272Google ScholarThere is no corresponding record for this reference.
- 358Pedersen, M. N.; Hedegård, E. D.; Olsen, J. M. H.; Kauczor, J.; Norman, P.; Kongsted, J. Damped Response Theory in Combination with Polarizable Environments: The Polarizable Embedding Complex Polarization Propagator Method J. Chem. Theory Comput. 2014, 10 (3) 1164– 1171 DOI: 10.1021/ct400946kGoogle Scholar358Damped Response Theory in Combination with Polarizable Environments: The Polarizable Embedding Complex Polarization Propagator MethodPedersen, Morten N.; Hedegaard, Erik D.; Olsen, Jogvan Magnus H.; Kauczor, Joanna; Norman, Patrick; Kongsted, JacobJournal of Chemical Theory and Computation (2014), 10 (3), 1164-1171CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present a combination of the polarizable embedding (PE) scheme with the complex polarization propagator (CPP) method with the aim of calcg. response properties including relaxation for large and complex systems. This new approach, termed PE-CPP, will benefit from the highly advanced description of the environmental electrostatic potential and polarization in the PE method as well as the treatment of near-resonant effects in the CPP approach. The PE-CPP model has been implemented in a Kohn-Sham d. functional theory approach, and we present pilot calcns. exemplifying the implementation for the UV/vis and carbon K-edge X-ray absorption spectra of the protein plastocyanin. Furthermore, tech. details assocd. with a PE-CPP calcn. are discussed.
- 359Olsen, J. M.; Aidas, K.; Kongsted, J. Excited States in Solution through Polarizable Embedding J. Chem. Theory Comput. 2010, 6 (12) 3721– 3734 DOI: 10.1021/ct1003803Google Scholar359Excited States in Solution through Polarizable EmbeddingOlsen, Jogvan Magnus; Aidas, Kestutis; Kongsted, JacobJournal of Chemical Theory and Computation (2010), 6 (12), 3721-3734CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present theory and implementation of an advanced quantum mechanics/mol. mechanics (QM/MM) approach using a fully self-consistent polarizable embedding (PE) scheme. It is a polarizable layered model designed for effective yet accurate inclusion of an anisotropic medium in a quantum mech. calcn. The polarizable embedding potential is described by an atomistic representation including terms up to localized octupoles and anisotropic polarizabilities. It is generally applicable to any quantum chem. description but is here implemented for the case of Kohn-Sham d. functional theory which we denote the PE-DFT method. It has been implemented in combination with time-dependent quantum mech. linear and nonlinear response techniques, thus allowing for assessment of electronic excitation processes and dynamic ground- and excited-state mol. properties using a nonequil. formulation of the environmental response. In our formulation of polarizable embedding we explicitly take into account the full self-consistent many-body environmental response from both ground and excited states. The PE-DFT method can be applied to any mol. system, e.g., proteins, nanoparticles and solute-solvent systems. Here, we present numerical examples of solvent shifts and excited-state properties related to a set of org. mols. in aq. soln.
- 360Sjöqvist, J.; Linares, M.; Mikkelsen, K. V.; Norman, P. QM/MM-MD Simulations of Conjugated Polyelectrolytes: A Study on Luminescent Conjugated Oligothiophenes for Use as Bio-Physical Probes J. Phys. Chem. A 2014, 118 (19) 3419– 3428 DOI: 10.1021/jp5009835Google ScholarThere is no corresponding record for this reference.
- 361Avila Ferrer, F. J.; Cerezo, J.; Stendardo, E.; Improta, R.; Santoro, F. Insights for an Accurate Comparison of Computational Data to Experimental Absorption and Emission Spectra: Beyond the Vertical Transition Approximation J. Chem. Theory Comput. 2013, 9 (4) 2072– 2082 DOI: 10.1021/ct301107mGoogle Scholar361Insights for an Accurate Comparison of Computational Data to Experimental Absorption and Emission Spectra: Beyond the Vertical Transition ApproximationAvila Ferrer, Francisco J.; Cerezo, Javier; Stendardo, Emiliano; Improta, Roberto; Santoro, FabrizioJournal of Chemical Theory and Computation (2013), 9 (4), 2072-2082CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The authors carefully study the relation between computed data and exptl. electronic spectra. To that end, the authors compare both vertical transition energies, EV, and characteristic frequencies of the spectrum like the max., νmax, and the center of gravity, M1, taking advantage of an anal. expression of M1 in terms of the parameters of the initial- and final-state potential energy surfaces. After pointing out that, for an accurate comparison, exptl. spectra should be preliminarily mapped from wavelength to frequency domain and transformed to normalized lineshapes, the authors simulate the absorption and emission spectra of several prototypical chromophores, obtaining lineshapes in very good agreement with exptl. data. Results indicate that the customary comparison of exptl. νmax and computational EV, without taking into account vibrational effects, is not an adequate measure of the performance of an electronic method. In fact, it introduces systematic errors that, in the studied systems, are ∼0.1-0.3 eV, i.e., values comparable to the expected accuracy of the most accurate computational methods. On the contrary, a comparison of exptl. and computed M1 and/or 0-0 transition frequencies provides more robust results. Some rules of thumbs probably help rationalize which kind of correction 1 should expect when comparing EV, M1, and νmax.
- 362Avila Ferrer, F. J.; Cerezo, J.; Soto, J.; Improta, R.; Santoro, F. First-Principle Computation of Absorption and Fluorescence Spectra in Solution Accounting for Vibronic Structure, Temperature Effects and Solvent Inhomogenous Broadening Comput. Theor. Chem. 2014, 1040–1041, 328– 337 DOI: 10.1016/j.comptc.2014.03.003Google Scholar362First-principle computation of absorption and fluorescence spectra in solution accounting for vibronic structure, temperature effects and solvent inhomogeneous broadeningAvila Ferrer, Francisco Jose; Cerezo, Javier; Soto, Juan; Improta, Roberto; Santoro, FabrizioComputational & Theoretical Chemistry (2014), 1040-1041 (), 328-337CODEN: CTCOA5; ISSN:2210-271X. (Elsevier B.V.)We compute the line shape of absorption and emission electronic spectra of two different dyes, coumarin C153 and N-methyl-6-quinolinium betaine accounting for the vibronic structure, temp. effects and polar solvent inhomogeneous broadening, without using any phenomenol. parameter. We exploit a no. of recent developments including a time-dependent (TD) approach to the computation of vibronic spectra that provides fully converged line shapes at finite temp. accounting for both Duschinsky and Herzberg-Teller effects, and the state-specific (SS) implementation of Polarizable Continuum Model (PCM). This latter is adopted to compute the solvent reorganization energy connected to inhomogeneous broadening. We compute the absorption and fluorescence spectra in the gas-phase, non-polar and polar solvents analyzing the relative importance of different sources of broadening. To this end we investigate the performance of time-dependent d. functional theory, complete active space self consistent field (CASSCF), and complete active space 2nd-order perturbation theory (CASPT2) methods in the computation of inhomogeneous broadening.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 454 publications.
- Yongjiao Song, Wanqiu Wang, Rui Liu. Visualization Experiment of Ion Migration in Electrolytic Cell Based on Natural Indicator. Journal of Chemical Education 2025, Article ASAP.
- Wei-Qin Xu, Di-En Huang, Jiao Li, Yin-Li Dai, Xu Feng, An-Qi Pang, Ling Lin, Min-Chang Chen, Wen-Ting Liu, Jun-Xing Zhong. A Home Chemistry Experiment Based on Chinese Traditional Culture: Anthocyanins in Peking Opera Facial Makeup and Paper-Cutting. Journal of Chemical Education 2025, 102
(1)
, 372-378. https://doi.org/10.1021/acs.jchemed.4c01117
- Peiqing Yang, Nuno Basílio, Xiaojun Liao, Zhenzhen Xu, Olivier Dangles, Fernando Pina. Influence of Acylation by Hydroxycinnamic Acids on the Reversible and Irreversible Processes of Anthocyanins in Acidic to Basic Aqueous Solution. Journal of Agricultural and Food Chemistry 2024, 72
(46)
, 25955-25971. https://doi.org/10.1021/acs.jafc.4c06847
- Olivier Dangles. Anthocyanins as Natural Food Colorings: The Chemistry Behind and Challenges Still Ahead. Journal of Agricultural and Food Chemistry 2024, 72
(22)
, 12356-12372. https://doi.org/10.1021/acs.jafc.4c01050
- María Oyón-Ardoiz, Elvira Manjón, María Teresa Escribano-Bailón, Ignacio García-Estévez. Potential Use of Torulaspora delbrueckii As a New Source of Mannoproteins of Oenological Interest. Journal of Agricultural and Food Chemistry 2024, 72
(20)
, 11606-11616. https://doi.org/10.1021/acs.jafc.4c01001
- Bárbara Torres-Rochera, Elvira Manjón, Natércia F Brás, María Teresa Escribano-Bailón, Ignacio García-Estévez. Supramolecular Study of the Interactions between Malvidin-3-O-Glucoside and Wine Phenolic Compounds: Influence on Color. Journal of Agricultural and Food Chemistry 2024, 72
(4)
, 1894-1901. https://doi.org/10.1021/acs.jafc.2c08502
- Tobias B. Koch, Anna M. Gabler, Florian Biener, Johanna Kreißl, Oliver Frank, Corinna Dawid, Heiko Briesen. Investigating the Role of Odorant–Polymer Interactions in the Aroma Perception of Red Wine: A Density Functional Theory-Based Approach. Journal of Agricultural and Food Chemistry 2023, 71
(50)
, 20231-20242. https://doi.org/10.1021/acs.jafc.3c03443
- Qianyu Zhao, Hua Zhang, Haitian Zhao, Hongwei Zhu, Jia Liu, Bin Li, Minjie Li, Xin Yang. Construction of a Biomimetic Receptor Based on Hydrophilic Multifunctional Monomer Covalent Organic Framework Molecularly Imprinted Polymers for Molecular Recognition of Cyanidin-3-O-Glucoside. Journal of Agricultural and Food Chemistry 2023, 71
(46)
, 18024-18036. https://doi.org/10.1021/acs.jafc.3c04391
- Judith Bijlsma, Wouter J. C. de Bruijn, Jamie Koppelaar, Mark G. Sanders, Krassimir P. Velikov, Jean-Paul Vincken. Interactions of Natural Flavones with Iron Are Affected by 7-O-Glycosylation, but Not by Additional 6″-O-Acylation. ACS Food Science & Technology 2023, 3
(6)
, 1111-1121. https://doi.org/10.1021/acsfoodscitech.3c00112
- Bohan Ma, Bin Dai, Xinghua Zhao, Mengge Zhao, Xingbin Yang, Langjun Cui, Hongjun Shao. Color-Stabilizing Effects and Mechanisms of Schisandra Lignans on the Anthocyanins Enriched with Petunidin-3-O-glucoside. ACS Food Science & Technology 2023, 3
(4)
, 717-728. https://doi.org/10.1021/acsfoodscitech.2c00435
- Rafael Grande, Riikka Räisänen, Jinze Dou, Satu Rajala, Kiia Malinen, Paula A. Nousiainen, Monika Österberg. In Situ Adsorption of Red Onion (Allium cepa) Natural Dye on Cellulose Model Films and Fabrics Exploiting Chitosan as a Natural Mordant. ACS Omega 2023, 8
(6)
, 5451-5463. https://doi.org/10.1021/acsomega.2c06650
- André Seco, Nuno Basílio, Natércia F. Brás, Kumi Yoshida, Tadao Kondo, Kin-ichi Oyama, Fernando Pina. Intermolecular Copigmentation Between Delphinidin 3-O-Glucoside and Chlorogenic Acid: Taking into Account the Existence of Neutral and Negatively Charged Forms of the Copigment. Journal of Agricultural and Food Chemistry 2022, 70
(36)
, 11391-11400. https://doi.org/10.1021/acs.jafc.2c03879
- Ambrósio Camuenho, André Seco, A. Jorge Parola, Nuno Basílio, Fernando Pina. Intermolecular Copigmentation of Malvidin-3-O-glucoside with Caffeine in Water: The Effect of the Copigment on the pH-Dependent Reversible and Irreversible Processes. ACS Omega 2022, 7
(29)
, 25502-25509. https://doi.org/10.1021/acsomega.2c02571
- Ellia H. La, M. Monica Giusti. Ultraviolet–Visible Excitation of cis- and trans-p-Coumaric Acylated Delphinidins and Their Resulting Photochromic Characteristics. ACS Food Science & Technology 2022, 2
(5)
, 878-887. https://doi.org/10.1021/acsfoodscitech.2c00028
- Rattapol Meelapsom, Waranphat Rattanakaroonjit, Akarapong Prakobkij, Nutthaporn Malahom, Saksri Supasorn, Sukhum Ruangchai, Purim Jarujamrus. Smartphone-Assisted Colorimetric Determination of Iron Ions in Water by Using Anthocyanin from Ruellia tuberosa L. as a Green Indicator and Application for Hands-on Experiment Kit. Journal of Chemical Education 2022, 99
(4)
, 1660-1671. https://doi.org/10.1021/acs.jchemed.1c01120
- Cláudia Novais, Adriana K. Molina, Rui M. V. Abreu, Celestino Santo-Buelga, Isabel C. F. R. Ferreira, Carla Pereira, Lillian Barros. Natural Food Colorants and Preservatives: A Review, a Demand, and a Challenge. Journal of Agricultural and Food Chemistry 2022, 70
(9)
, 2789-2805. https://doi.org/10.1021/acs.jafc.1c07533
- Lucia Panzella, Alessandra Napolitano. Condensed Tannins, a Viable Solution To Meet the Need for Sustainable and Effective Multifunctionality in Food Packaging: Structure, Sources, and Properties. Journal of Agricultural and Food Chemistry 2022, 70
(3)
, 751-758. https://doi.org/10.1021/acs.jafc.1c07229
- Luis Cruz, Nuno Basílio, Nuno Mateus, Victor de Freitas, Fernando Pina. Natural and Synthetic Flavylium-Based Dyes: The Chemistry Behind the Color. Chemical Reviews 2022, 122
(1)
, 1416-1481. https://doi.org/10.1021/acs.chemrev.1c00399
- Fabian Weber. Noncovalent Polyphenol–Macromolecule Interactions and Their Effects on the Sensory Properties of Foods. Journal of Agricultural and Food Chemistry 2022, 70
(1)
, 72-78. https://doi.org/10.1021/acs.jafc.1c05873
- Xu Zhao, Xinke Zhang, Xiaoming He, Changqing Duan, Fei He. Acetylation of Malvidin-3-O-glucoside Impedes Intermolecular Copigmentation: Experimental and Theoretical Investigations. Journal of Agricultural and Food Chemistry 2021, 69
(27)
, 7733-7741. https://doi.org/10.1021/acs.jafc.1c02378
- Luís Cruz, Juan Correa, Nuno Mateus, Victor de Freitas, Maun H. Tawara, Eduardo Fernandez-Megia. Dendrimers as Color-Stabilizers of Pyranoanthocyanins: The Dye Concentration Governs the Host–Guest Interaction Mechanisms. ACS Applied Polymer Materials 2021, 3
(3)
, 1457-1464. https://doi.org/10.1021/acsapm.0c01321
- Joana Oliveira, Joana Azevedo, Natércia Teixeira, Paula Araújo, Victor de Freitas, Nuno Basílio, Fernando Pina. On the Limits of Anthocyanins Co-Pigmentation Models and Respective Equations. Journal of Agricultural and Food Chemistry 2021, 69
(4)
, 1359-1367. https://doi.org/10.1021/acs.jafc.0c05954
- Laura Estévez, Marta Queizán, Ricardo A. Mosquera, Lucia Guidi, Ermes Lo Piccolo, Marco Landi. First Characterization of the Formation of Anthocyanin–Ge and Anthocyanin–B Complexes through UV–Vis Spectroscopy and Density Functional Theory Quantum Chemical Calculations. Journal of Agricultural and Food Chemistry 2021, 69
(4)
, 1272-1282. https://doi.org/10.1021/acs.jafc.0c06827
- Cátia I. Sampaio, Luís F. Sousa, Alice M. Dias. Separation of Anthocyaninic and Nonanthocyaninic Flavonoids by Liquid–Liquid Extraction Based on Their Acid–Base Properties: A Green Chemistry Approach. Journal of Chemical Education 2020, 97
(12)
, 4533-4539. https://doi.org/10.1021/acs.jchemed.0c00139
- Savanah G. Reeves, Arpad Somogyi, Wayne E. Zeller, Theresa A. Ramelot, Kelly C. Wrighton, Ann E. Hagerman. Proanthocyanidin Structural Details Revealed by Ultrahigh Resolution FT-ICR MALDI-Mass Spectrometry, 1H–13C HSQC NMR, and Thiolysis-HPLC–DAD. Journal of Agricultural and Food Chemistry 2020, 68
(47)
, 14038-14048. https://doi.org/10.1021/acs.jafc.0c04877
- Daisuke Kajiya. Demonstrating Purple Color Development to Students by Showing the Highly Visual Effects of Aluminum Ions and pH on Aqueous Anthocyanin Solutions. Journal of Chemical Education 2020, 97
(11)
, 4084-4090. https://doi.org/10.1021/acs.jchemed.0c00476
- Hiléia K. S. Souza, Nuno Mateus, Victor de Freitas, Maria Pilar Gonçalves, Luís Cruz. Chemical/Color Stability and Rheological Properties of Cyanidin-3-Glucoside in Deep Eutectic Solvents as a Gateway to Design Task-Specific Bioactive Compounds. ACS Sustainable Chemistry & Engineering 2020, 8
(43)
, 16184-16196. https://doi.org/10.1021/acssuschemeng.0c04839
- Michal Langer, Markéta Paloncýová, Miroslav Medved’, Michal Otyepka. Molecular Fluorophores Self-Organize into C-Dot Seeds and Incorporate into C-Dot Structures. The Journal of Physical Chemistry Letters 2020, 11
(19)
, 8252-8258. https://doi.org/10.1021/acs.jpclett.0c01873
- Kenneth R. Sims, Jr., Brian He, Hyun Koo, Danielle S.W. Benoit. Electrostatic Interactions Enable Nanoparticle Delivery of the Flavonoid Myricetin. ACS Omega 2020, 5
(22)
, 12649-12659. https://doi.org/10.1021/acsomega.9b04101
- Federica Moccia, Sarai Agustin-Salazar, Anna-Lisa Berg, Brunella Setaro, Raffaella Micillo, Elio Pizzo, Fabian Weber, Nohemi Gamez-Meza, Andreas Schieber, Pierfrancesco Cerruti, Lucia Panzella, Alessandra Napolitano. Pecan (Carya illinoinensis (Wagenh.) K. Koch) Nut Shell as an Accessible Polyphenol Source for Active Packaging and Food Colorant Stabilization. ACS Sustainable Chemistry & Engineering 2020, 8
(17)
, 6700-6712. https://doi.org/10.1021/acssuschemeng.0c00356
- Woojin Yang, Junhyuk Jo, Hyeongyeol Oh, Hohjai Lee, Won-jin Chung, Jiwon Seo. Peptoid Helix Displaying Flavone and Porphyrin: Synthesis and Intramolecular Energy Transfer. The Journal of Organic Chemistry 2020, 85
(3)
, 1392-1400. https://doi.org/10.1021/acs.joc.9b02358
- Tomomi Ujihara, Nobuyuki Hayashi. Mechanism of Copigmentation of Monoglucosylrutin with Caffeine. Journal of Agricultural and Food Chemistry 2020, 68
(1)
, 323-331. https://doi.org/10.1021/acs.jafc.9b06235
- Andrzej Sienkiewicz, Iwona Rusinek, Anna Siatecka, Sonia Losada-Barreiro. Flower Color Change Demonstration as a Visualization of Potential Harmful Effects Associated with Ammonia Gas on Living Organisms. Journal of Chemical Education 2019, 96
(9)
, 1982-1987. https://doi.org/10.1021/acs.jchemed.9b00001
- Chen Tan, Mohammad Arshadi, Michelle C. Lee, Mary Godec, Morteza Azizi, Bing Yan, Hamed Eskandarloo, Ted W. Deisenroth, Rupa Hiremath Darji, Toan Van Pho, Alireza Abbaspourrad. A Robust Aqueous Core–Shell–Shell Coconut-like Nanostructure for Stimuli-Responsive Delivery of Hydrophilic Cargo. ACS Nano 2019, 13
(8)
, 9016-9027. https://doi.org/10.1021/acsnano.9b03049
- Johan Mendoza, Nuno Basílio, Victor de Freitas, Fernando Pina. New Procedure To Calculate All Equilibrium Constants in Flavylium Compounds: Application to the Copigmentation of Anthocyanins. ACS Omega 2019, 4
(7)
, 12058-12070. https://doi.org/10.1021/acsomega.9b01066
- Cristina Alcalde-Eon, Claudia Pérez-Mestre, Rebeca Ferreras-Charro, Francisco J. Rivero, Francisco J. Heredia, María Teresa Escribano-Bailón. Addition of Mannoproteins and/or Seeds during Winemaking and Their Effects on Pigment Composition and Color Stability. Journal of Agricultural and Food Chemistry 2019, 67
(14)
, 4031-4042. https://doi.org/10.1021/acs.jafc.8b06922
- Vijayakumar
R. Vishnu, Raveendran S. Renjith, Archana Mukherjee, Shirly Raichal Anil, Janardanan Sreekumar, Alummoottil N. Jyothi. Comparative Study on the Chemical Structure and In Vitro Antiproliferative Activity of Anthocyanins in Purple Root Tubers and Leaves of Sweet Potato (Ipomoea batatas). Journal of Agricultural and Food Chemistry 2019, 67
(9)
, 2467-2475. https://doi.org/10.1021/acs.jafc.8b05473
- Tomomi Ujihara, Nobuyuki Hayashi. Complex Structures of Monoglucosylrutin with ent-Gallocatechin-3-O-gallate and Epigallocatechin-3-O-gallate in Aqueous Solutions and the Mechanism of Color Change Induced by Complexation. Journal of Natural Products 2019, 82
(1)
, 2-8. https://doi.org/10.1021/acs.jnatprod.7b00817
- Imene Bayach, Anthony D’Aleó, Patrick Trouillas. Tuning Optical Properties of Chalcone Derivatives: A Computational Study. The Journal of Physical Chemistry A 2019, 123
(1)
, 194-201. https://doi.org/10.1021/acs.jpca.8b08529
- A. Alejo-Armijo, Livia Corici, Liliana Cseh, Diana Aparaschivei, Artur J. Moro, A. Jorge Parola, João C. Lima, Fernando Pina. Achieving Complexity at the Bottom. 2,6-Bis(arylidene)cyclohexanones and Anthocyanins: The Same General Multistate of Species. ACS Omega 2018, 3
(12)
, 17853-17862. https://doi.org/10.1021/acsomega.8b02745
- Laura
A. Chatham, Leslie West, Mark A. Berhow, Karl E. Vermillion, John A. Juvik. Unique Flavanol-Anthocyanin Condensed Forms in Apache Red Purple Corn. Journal of Agricultural and Food Chemistry 2018, 66
(41)
, 10844-10854. https://doi.org/10.1021/acs.jafc.8b04723
- Paul M. Rose, Victoria Cantrill, Meryem Benohoud, Alenka Tidder, Christopher M. Rayner, Richard S. Blackburn. Application of Anthocyanins from Blackcurrant (Ribes nigrum L.) Fruit Waste as Renewable Hair Dyes. Journal of Agricultural and Food Chemistry 2018, 66
(26)
, 6790-6798. https://doi.org/10.1021/acs.jafc.8b01044
- Paula Araújo, Nuno Basílio, Ana Fernandes, Nuno Mateus, Victor de Freitas, Fernando Pina, Joana Oliveira. Impact of Lignosulfonates on the Thermodynamic and Kinetic Parameters of Malvidin-3-O-glucoside in Aqueous Solutions. Journal of Agricultural and Food Chemistry 2018, 66
(25)
, 6382-6387. https://doi.org/10.1021/acs.jafc.8b02273
- Johan Mendoza, Nuno Basílio, Fernando Pina, Tadao Kondo, Kumi Yoshida. Rationalizing the Color in Heavenly Blue Anthocyanin: A Complete Kinetic and Thermodynamic Study. The Journal of Physical Chemistry B 2018, 122
(19)
, 4982-4992. https://doi.org/10.1021/acs.jpcb.8b01136
- Chen Tan, Giovana B. Celli, Michelle Lee, Jonathan Licker, Alireza Abbaspourrad. Polyelectrolyte Complex Inclusive Biohybrid Microgels for Tailoring Delivery of Copigmented Anthocyanins. Biomacromolecules 2018, 19
(5)
, 1517-1527. https://doi.org/10.1021/acs.biomac.8b00352
- Alberto Fabrizio and Clémence Corminboeuf . How do London Dispersion Interactions Impact the Photochemical Processes of Molecular Switches?. The Journal of Physical Chemistry Letters 2018, 9
(3)
, 464-470. https://doi.org/10.1021/acs.jpclett.7b03316
- Cassio Pacheco da Silva, Renan Moraes Pioli, Liu Liu, Shasha Zheng, Mengjiao Zhang, Gustavo Thalmer de Medeiros Silva, Vânia Maria Teixeira Carneiro, and Frank H. Quina . Improved Synthesis of Analogues of Red Wine Pyranoanthocyanin Pigments. ACS Omega 2018, 3
(1)
, 954-960. https://doi.org/10.1021/acsomega.7b01955
- Kateřina Holá, Mária Sudolská, Sergii Kalytchuk, Dana Nachtigallová, Andrey L. Rogach, Michal Otyepka, and Radek Zbořil . Graphitic Nitrogen Triggers Red Fluorescence in Carbon Dots. ACS Nano 2017, 11
(12)
, 12402-12410. https://doi.org/10.1021/acsnano.7b06399
- Yebeen Kang, Eunmi Koh. Chlorogenic acid is more effective than aluminum ion for protecting anthocyanins in aqueous aronia solution against LED light. Journal of Food Composition and Analysis 2025, 143 , 107557. https://doi.org/10.1016/j.jfca.2025.107557
- Sabrina Idir, Sabiha Achat, Luis Cruz, Olivier Dangles. Anthocyanin-rich extracts: Susceptibility to color loss by hydration and thermal degradation, influence of metal ions and endogenous copigments. Food Chemistry 2025, 481 , 144004. https://doi.org/10.1016/j.foodchem.2025.144004
- Mariana C. Leal-Alcazar, Frida Bautista-Palestina, María del R. Rocha-Pizaña, Luis Mojica, Alan Javier Hernández-Álvarez, Diego A. Luna-Vital. Extraction, stabilization, and health application of betalains: An update. Food Chemistry 2025, 481 , 144011. https://doi.org/10.1016/j.foodchem.2025.144011
- Giroon Ijod, Nur Izzati Mohamed Nawawi, Mohammed S. Qoms, Mohammad Rashedi Ismail Fitry, Muhamad Hafiz Abd Rahim, Dimitris Charalampopoulos, Rabiha Sulaiman, Noranizan Mohd Adzahan, Ezzat Mohamad Azman. Synergistic effects of intermolecular copigmentation and high-pressure processing on stabilizing mangosteen pericarp anthocyanins. Food Chemistry 2025, 480 , 143888. https://doi.org/10.1016/j.foodchem.2025.143888
- Caiyun Liu, Lulu Wu, Zengshuai Zhang, Ziying Li, Mario Prejanò, Tiziana Marino, Yongsheng Tao, Yunkui Li. Copigmentation effect and mechanism of α-cyclodextrin on wine color quality and stability: Combining dynamics, thermodynamics, structural characterization and quantum mechanics. Food Hydrocolloids 2025, 163 , 111068. https://doi.org/10.1016/j.foodhyd.2025.111068
- Oscar Zannou, Ilkay Koca, Reza Tahergorabi, Salam A. Ibrahim. Deep eutectic solvent as an alternative, green medium for enhancing the stability of blackberry (Rubus spp) pigments. Industrial Crops and Products 2025, 228 , 120866. https://doi.org/10.1016/j.indcrop.2025.120866
- Zixuan Li, Zhipeng Gao, Donglin Su, Zhipeng Su, Yanjiao Fu, Wenbin Xiao, Fuhua Fu, Gaoyang Li, Lvhong Huang, Jiajing Guo, Yang Shan. Facile and green synthesis of nanocarriers from chondroitin sulfate-based and whey protein isolate with improved stability and biocompatibility for anthocyanins protection. Food Chemistry 2025, 476 , 143449. https://doi.org/10.1016/j.foodchem.2025.143449
- Jiaheng Lyu, Jiming Li, Wenguang Jiang, Tingting Liu, Yan Xu, Ke Tang. Copigmentation effects of different phenolics on color stability of three basic anthocyanins in wines: Chromaticity, thermodynamics and molecular dynamics simulation. Food Chemistry 2025, 476 , 143499. https://doi.org/10.1016/j.foodchem.2025.143499
- Gerui Ren, Xinpei Cai, Ying He, Lei Liu, Ruiqi Hu, Jiqi Wang, Qingbo Jiao, Yingjie Wu, Jiacheng Liu, Ying Huang, Min Huang, Hujun Xie, Kejun Cheng. Ferulic acid-regulated anthocyanin-based intelligent film: A promising strategy to improve stability, sensitivity, and biological activity. International Journal of Biological Macromolecules 2025, 305 , 141237. https://doi.org/10.1016/j.ijbiomac.2025.141237
- Nadja Ulmann, Johnny Hioe, Didier Touraud, Eva Müller, Dominik Horinek, Werner Kunz. Sustainable polyphenolate assisted hydrotropic solubilization of riboflavin. Journal of Molecular Liquids 2025, 426 , 127465. https://doi.org/10.1016/j.molliq.2025.127465
- S. Rincon, H. Murray, M. Gössinger, C. Ginies, P. Goupy, C. Dufour, O. Dangles, C. Le Bourvellec. Characterisation of phenolic compounds and polysaccharides in strawberry: Cultivar and harvest effects and their correlation with nectar colour stability. Food Chemistry 2025, 473 , 143112. https://doi.org/10.1016/j.foodchem.2025.143112
- Fangfang Li, Quancai Sun, Long Chen, Ruojie Zhang, Zipei Zhang. Unlocking the health potential of anthocyanins: a structural insight into their varied biological effects. Critical Reviews in Food Science and Nutrition 2025, 65
(11)
, 2134-2154. https://doi.org/10.1080/10408398.2024.2328176
- Aleksandra Popowska, Joanna Oracz. Influence of Copigmentation and Encapsulation on Stability and Antioxidant Activity of Anthocyanins from Blue and Pink Cornflower (Centaurea cyanus L.) Flowers. Molecules 2025, 30
(7)
, 1467. https://doi.org/10.3390/molecules30071467
- Yuan Yuan, Caochuang Fang, Chaohan Li, Jiaqi You, Kun Ma. Identification and Evaluation of Colour Change in Rosemary and Biluochun Tea Infusions. Metabolites 2025, 15
(4)
, 265. https://doi.org/10.3390/metabo15040265
- Francisco Chamizo-González, Francisco J. Heredia, María Fernanda López-Molina, Francisco J. Rodríguez-Pulido, M. Lourdes González-Miret, Belén Gordillo. Theoretical Prediction of the Color Expression of Malvidin 3-Glucoside by in silico Tristimulus Colorimetry. Effects of Structure Conformational Changes and Molecular Interactions. Applied Sciences 2025, 15
(8)
, 4238. https://doi.org/10.3390/app15084238
- Xuechen Yao, Haoen Cai, Jiayi Kou, Yunxue Xie, Jin Li, Penghui Zhou, Fei He, Changqing Duan, Qiuhong Pan, Mengyao Qi, Yibin Lan. Dual-temperature dual-state fermentation: A novel approach to improve aroma and color characteristics of Marselan wines. Food Chemistry: X 2025, 27 , 102447. https://doi.org/10.1016/j.fochx.2025.102447
- Damla Ezgi Uzun, Tugce Ceyhan, Merve Tomas, Esra Capanoglu. Recent advances in improving anthocyanin stability in black carrots. Critical Reviews in Food Science and Nutrition 2025, 3 , 1-23. https://doi.org/10.1080/10408398.2025.2469774
- Shihan Bao, Jianing Li, Jiaqi Wang, Tian Lan, Mengyuan Wei, Xiangyu Sun, Yulin Fang, Tingting Ma. How Does Nature Create the Painting “Gradient Coloration of ‘Manicure Finger’ Grape”? Integrated Omics Unveil the Pigments Basis and Metabolism Networks of Its Formation. Food Frontiers 2025, 6
(2)
, 921-939. https://doi.org/10.1002/fft2.513
- Shaohua Zeng, Shuang Lin, Rong Jiang, Jinrong Wei, Ying Wang. Biotechnology Advances in Natural Food Colorant Acylated Anthocyanin Production. Food Frontiers 2025, 6
(2)
, 698-715. https://doi.org/10.1002/fft2.527
- Shiguang Chen, Zhiling Li, Dan Ren, Xiyu Wu, Dan Xu. Improved sensitivity of freshness indicator based on purple sweet potato anthocyanins through pH optimization and its application in flesh food monitoring during logistics. Innovative Food Science & Emerging Technologies 2025, 100 , 103929. https://doi.org/10.1016/j.ifset.2025.103929
- Jingliang Zhu, Yaqian Geng, Yuan Huang, Chen Ma, Li Dong, Fang Chen, Xiaosong Hu, Lingjun Ma, Junfu Ji. Emerging Novel Processing Technologies Towards the Stabilization of Anthocyanins. Food Reviews International 2025, 41
(2)
, 439-468. https://doi.org/10.1080/87559129.2024.2403525
- Bojian Chen, Lirong Xiang, Danyue Zhao, Zunying Liu, Fei Jia. Unraveling Nature’s Color Palette: The Chemistry, Biosynthesis and Applications in Health Promotion of Anthocyanins—A Comprehensive Review. Food Reviews International 2025, 41
(2)
, 491-520. https://doi.org/10.1080/87559129.2024.2404471
- Xinjie Li, Fan Wang, Na Ta, Jinyong Huang. The compositions, characteristics, health benefits and applications of anthocyanins in Brassica crops. Frontiers in Plant Science 2025, 16 https://doi.org/10.3389/fpls.2025.1544099
- María Fernanda López-Molina, Francisco J. Rodríguez-Pulido, Ana Belén Mora-Garrido, M. Lourdes González-Miret, Francisco J. Heredia. New approaches for screening grape seed peptides as colourimetric modulators by malvidin-3-O-glucoside stabilisation. Food Chemistry 2025, 464 , 141708. https://doi.org/10.1016/j.foodchem.2024.141708
- Zhiying Li, Jinlong Tian, Qilin Tian, Zhihuan Zang, Yumeng Wang, Qiao Jiang, Yi Chen, Baoru Yang, Shufang Yang, Yiyun Yang, Bin Li. Improved uptake of anthocyanins-loaded nanoparticles based on phenolic acid-grafted zein and lecithin. Food Chemistry 2025, 466 , 142235. https://doi.org/10.1016/j.foodchem.2024.142235
- Ruirui Li, Dongyue Yang, Zhiyu Li, Xiaohong Tang, Ke Zhong, Yan Ding, Xiaomei Han, Xueqiang Guan, Yuxia Sun. The effect of stem contact fermentation on the quality of Cabernet Sauvignon and Merlot wines from Yantai, China. Food Bioscience 2025, 64 , 105872. https://doi.org/10.1016/j.fbio.2025.105872
- Katarina Delic, Claire Payan, Viktoriya Aleksovych, A. Jouin, A. Vignault, Kleopatra Chira, Michael Jourdes, Pierre-Louis Teissedre. Wine Phenolic Compounds: Chemistry and Biological Properties. 2025, 713-759. https://doi.org/10.1007/978-3-031-38663-3_218
- Thomas Owen Hay, Melissa A. Fitzgerald, Joseph Robert Nastasi. Systematic application of UPLC-Q-ToF-MS/MS coupled with chemometrics for the identification of natural food pigments from Davidson plum and native currant. Food Chemistry: X 2025, 25 , 102072. https://doi.org/10.1016/j.fochx.2024.102072
- Niloofar Moshfegh, Mehrdad Niakousary, Seyed Mohammad Hashem Hosseini, Seyed Mohammad Mazloomi, Azam Abbasi. Effect of maltodextrin and Persian gum as wall materials and tannic acid as copigment on some properties of encapsulated sour cherry anthocyanin microcapsules. Food Chemistry 2025, 463 , 141165. https://doi.org/10.1016/j.foodchem.2024.141165
- Xiangxin Gu, Yaqiong Liu, Ran Suo, Qingquan Yu, Churan Xue, Jie Wang, Wenxiu Wang, Haiqi Wang, Yan Qiao. Effects of different low-temperature maceration times on the chemical and sensory characteristics of Syrah wine. Food Chemistry 2025, 463 , 141230. https://doi.org/10.1016/j.foodchem.2024.141230
- Sarvpreet Singh, Nitisha Sendri, Bhanu Sharma, Pramod Kumar, Avisha Sharma, Narendra Vijay Tirpude, Rituraj Purohit, Pamita Bhandari. Copigmentation effect on red cabbage anthocyanins, investigation of their cellular viability and interaction mechanism. Food Research International 2025, 200 , 115427. https://doi.org/10.1016/j.foodres.2024.115427
- Yuxi Liu, Rico F. Tabor, Piotr Pawliszak, David A. Beattie, Marta Krasowska, Benjamin W. Muir, San H. Thang, Chris Ritchie. Multi-stimuli-responsive polymers enabled by bio-inspired dynamic equilibria of flavylium chemistry. Chemical Science 2025, 122 https://doi.org/10.1039/D5SC00977D
- Shaojun Zuo, Tongtong Li, Tong Chen, Jianing Li, Xinyou Liu. Degradation of Oil Paint Coating Based on Wood Under the Combined Effect of UV Light and Heat. Forests 2025, 16
(1)
, 22. https://doi.org/10.3390/f16010022
- Lucas M. O. S. Martins, Gustavo T. M. Silva, Lucas F. S. Hess, Alexandre B. Barbosa, Claudia Turro, Mauricio S. Baptista, Frank H. Quina. Heavy atom effects on synthetic pyranoanthocyanin analogues. Photochemistry and Photobiology 2024, https://doi.org/10.1111/php.14058
- Tisong Liang, Pu Jing, Jian He. Nano techniques: an updated review focused on anthocyanin stability. Critical Reviews in Food Science and Nutrition 2024, 64
(32)
, 11985-12008. https://doi.org/10.1080/10408398.2023.2245893
- Yun Wang, David Julian McClements, Long Chen, Xinwen Peng, Zhenlin Xu, Man Meng, Hangyan Ji, Chaohui Zhi, Lei Ye, Jianwei Zhao, Zhengyu Jin. Progress on molecular modification and functional applications of anthocyanins. Critical Reviews in Food Science and Nutrition 2024, 64
(31)
, 11409-11427. https://doi.org/10.1080/10408398.2023.2238063
- Nitisha Sendri, Pamita Bhandari. Anthocyanins: a comprehensive review on biosynthesis, structural diversity, and industrial applications. Phytochemistry Reviews 2024, 23
(6)
, 1913-1974. https://doi.org/10.1007/s11101-024-09945-9
- Ana Beatriz Neves Martins, Ellen Cristina Quirino Lacerda, Daniel Perrone, Mariana Monteiro. Heat-extracted jaboticaba juice blended with apple juice has good sensory acceptance and the addition of Zn2+ delays the degradation of its anthocyanins during storage at room temperature. Applied Food Research 2024, 4
(2)
, 100482. https://doi.org/10.1016/j.afres.2024.100482
- Sarah Otto, Marta Krasowska, Stephanie MacWilliams, David Beattie, Anton Blencowe. The solid-state stability of anthocyanins under various conditions and the implications for storage and shelf-life. Dyes and Pigments 2024, 231 , 112367. https://doi.org/10.1016/j.dyepig.2024.112367
- Hongkun Xue, Jianduo Zhao, Yu Wang, Zhangmeng Shi, Kaifang Xie, Xiaojun Liao, Jiaqi Tan. Factors affecting the stability of anthocyanins and strategies for improving their stability: A review. Food Chemistry: X 2024, 24 , 101883. https://doi.org/10.1016/j.fochx.2024.101883
- Min Jeong Kang, Ronald B. Pegg, William L. Kerr, M. Lenny Wells, Patrick J. Conner, Joon Hyuk Suh. Metabolomic analysis combined with machine learning algorithms enables the evaluation of postharvest pecan color stability. Food Chemistry 2024, 461 , 140814. https://doi.org/10.1016/j.foodchem.2024.140814
- Ana Rita Pereira, Virgínia Cruz Fernandes, Cristina Delerue-Matos, Victor de Freitas, Nuno Mateus, Joana Oliveira. Exploring acylated anthocyanin-based extracts as a natural alternative to synthetic food dyes: Stability and application insights. Food Chemistry 2024, 461 , 140945. https://doi.org/10.1016/j.foodchem.2024.140945
- Jiaxin Li, Yiwen Bao, Qiao Jiang, Bo Wen, Liang Wang, Ying He, Xu Si, Bin Li. Indicator-enhanced starch-based intelligent film for nondestructive monitoring of beef freshness: Different structural phenolic acids copigment anthocyanin. Journal of Food Engineering 2024, 383 , 112241. https://doi.org/10.1016/j.jfoodeng.2024.112241
- Ana S. Márquez-Rodríguez, Arasay Carrión, Felipe Trejo, Hilda E. Esparza-Ponce, José M. Nápoles-Duarte, María L. Ballinas-Casarrubias, Luis E. Fuentes-Cobas, Erika Salas, Juan P. Palomares-Báez, María E. Fuentes-Montero. Anthocyanins stabilization of Hibiscus sabdariffa extract with sepiolite: Analytical and reactive force fields approaches. Sustainable Chemistry and Pharmacy 2024, 42 , 101831. https://doi.org/10.1016/j.scp.2024.101831
- Mochamad Nur Jalil Aripin, Helen Julian, Tjokorde Walmiki Samadhi. Enhancing the Quality of Natural Food Colorants by Novel Extraction and Concentration Processes. Food Reviews International 2024, 40
(10)
, 3899-3937. https://doi.org/10.1080/87559129.2024.2374842
- Zhongzheng Zhang, Yongxi Ren, Xiaoming He, Xinke Zhang, Guangren Pei, Xu Zhao. Exploring the impact of substitution and conformational variations on the copigmentation ability of monomeric flavan-3-ols in wine. Food Research International 2024, 196 , 115032. https://doi.org/10.1016/j.foodres.2024.115032
- Yanli Ma, Lei Wen, Jinxiao Liu, Pengfei Du, Yaobo Liu, Peng Hu, Jianfang Cao, Weiting Wang. Enhanced pH-sensitive anthocyanin film based on chitosan quaternary ammonium salt: A promising colorimetric indicator for visual pork freshness monitoring. International Journal of Biological Macromolecules 2024, 279 , 135236. https://doi.org/10.1016/j.ijbiomac.2024.135236
- Xiaohui Di, Yaochang Li, Xinguang Qin, Qi Wang, Gang Liu. Investigating the effect of whey protein isolate:proanthocyanidin complex ratio on the stability and antioxidant capacity of Pickering emulsions. International Journal of Biological Macromolecules 2024, 279 , 135342. https://doi.org/10.1016/j.ijbiomac.2024.135342
- Guixin Tan, Jingjie Hou, Dekun Meng, Huajiang Zhang, Xiue Han, Hanyu Li, Zhongjiang Wang, Mohamed Ghamry, Ahmed M. Rayan. 3D printing cassava starch-ovalbumin intelligent labels: Co-pigmentation effects of gallic acid on anthocyanins. International Journal of Biological Macromolecules 2024, 281 , 135684. https://doi.org/10.1016/j.ijbiomac.2024.135684
- Xuebing Bai, Xinlong Chen, Xiaohan Li, Fangdai Tan, Faisal Eudes Sam, Yongsheng Tao. Wine polyphenol oxidation mechanism and the effects on wine quality: A review. Comprehensive Reviews in Food Science and Food Safety 2024, 23
(6)
https://doi.org/10.1111/1541-4337.70035
- Xuefu Zhou, Pangzhen Zhang, Danyang Ying, Xu Duan, Zhongxiang Fang. 3‐Deoxyanthocyanidins: Extraction, stability, and food applications. Comprehensive Reviews in Food Science and Food Safety 2024, 23
(6)
https://doi.org/10.1111/1541-4337.70064
- Nuo Wang, Xiu-Jun Li, Liang Wang, Bin Li, Jin-Long Tian. Design of a liposome casein hydrogel as an efficient front-end homeostatic anthocyanin loading system. International Journal of Biological Macromolecules 2024, 278 , 134928. https://doi.org/10.1016/j.ijbiomac.2024.134928
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Scheme 1
Scheme 1. Prototypical Copigmentation ComplexesaScheme a(A) Noncovalent association of a prototypical anthocyanin pigment and a prototypical flavonoid copigment (intermolecular copigmentation). (B) Prototypical acylated derivatives allowing copigmentation between the anthocyanin moiety and two phenolic acids covalently linked (intramolecular copigmentation).
Figure 1
Figure 1. Chemical structures of anthocyanidins (R3 = OH) and anthocyanins (R3 = glucose).
Figure 2
Figure 2. Chemical structures of pyranoanthocyanins (flavylium cation form).
Figure 3
Figure 3. Structural transformations of anthocyanins in acidic to neutral solution. Proton loss from C5-OH omitted for simplicity.
Figure 4
Figure 4. Structural transformations of the 3′,4′-dihydroxy-7-O-β-d-glucopyranosyloxyflavylium ion. (A) Fast formation of the quinonoid base. Spectral measurements immediately after addition of pigment to buffer, pH = 2.0 → pH = 6.0. (B) Slow accumulation of trans-chalcone (Ctrans). Spectral measurements on solutions equilibrated overnight, pH = 2.0 → pH = 4.0. Adapted with permission from ref 25. Copyright 2013 Elsevier Ltd.
Figure 5
Figure 5. Chemical structures of series of copigments.
Scheme 2
Scheme 2. π–π Stacking Interactions in Anthocyanins and Their ComplexesaScheme a(A) Intermolecular copigmentation, (B) self-association, (C) intramolecular copigmentation in acylated anthocyanins, (D) self-association of acylated anthocyanins, (E) intercalation in intermolecular copigmentation, and (F) copigmentation in metal–anthocyanin complexes.
Figure 6
Figure 6. Chemical structures of three nonmetallic pigment···copigment assemblies.
Figure 7
Figure 7. Chemical structures of two polyacylated anthocyanins.
Figure 8
Figure 8. Chemical structures of polyacylated anthocyanins.
Figure 9
Figure 9. Anthocyanin-flavone C- or O-glycoside pigments with a malonate bridge from leaves of (A) Oxalis triangularis and (B) Allium “Blue Perfume”.
Figure 10
Figure 10. Structure of commelinin. Blue: malonylawobanin (MA), yellow and orange: flavocommelin (FC), red: Mg2+(A). A side view of a left-handed stacking of two MA units that coordinate to different Mg ions. (B) A side view of a copigmentation of MA and FC in a right-handed stacking arrangement. (C) A side view of a left-handed stacking of two FC units. (D) A side view of commelinin. (E) A skew view of a copigmentation between MA and FC. (F) A skew view of the self-association of two FC units. Adapted with permission from ref 1. Copyright 2009 The Royal Society of Chemistry.
Figure 11
Figure 11. Chemical structure of the anthocyanin derived from Ceanothus papillosus petals.
Figure 12
Figure 12. (A) Structure of protocyanins from Centaurea cyanus (cornflower). (B) X-ray diffraction crystal structure, showing the two metal ions Fe3+ and Mg2+ in the center of the supramolecular assembly. Adapted with permission from ref 1. Copyright 2009 The Royal Society of Chemistry.
Figure 13
Figure 13. Copigmentation of malvin (malvidin 3,5-di-O-β-d-glucoside) by 7-O-sulfoquercetin. Reprinted with permission from ref 107. Copyright 2001 John Wiley & Sons.
Figure 14
Figure 14. pH dependence of the copigmentation hyperchromic shift. Spectroscopic monitoring at λMAX of free flavylium. Simulations from eq 2 using the following parameters: CP concentration = 10 mM, pK′h = 2.5, pKa = 4.0, rAHCP+ = 0.8, rACP = 0.9, rA = 0.7. Hydroxycinnamic acids (red line): KAHCP+ = 3 × 102 M–1, KACP = 2 × 102 M–1. Flavones and flavonols (green line): KAHCP+ = 3 × 103 M–1, KACP = 103 M–1.
Figure 15
Figure 15. Distribution diagram of anthocyanin forms in the presence of a copigment. Parameters: pK′h = 2.5, pKa = 4.0, pigment and copigment concentrations = 10 and 30 mM, respectively. (A) Hydroxycinnamic acid, KAHCP = 3 × 102, KACP = 2 × 102 M–1, pK′hCP (pH of a 1:1 mixture of colored and colorless forms) ≈ 3.5. (B) Flavone or flavonol, KAHCP = 3 × 102 M−1, KAHCP = 3 × 103 M–1, pK′hCP ≈ 5.2.
Figure 16
Figure 16. Morning glory (Pharbitis nil) anthocyanins (inspired from ref 135).
Figure 17
Figure 17. Distribution diagram for a prototypical acylated anthocyanin forms showing the strong influence of noncovalent dimerization. Parameters: pK′h = 3.0 (assuming partial protection against water addition owing to intramolecular copigmentation in monomer), pKa = 4.0, pigment concentration = 10 mM. Tentative values for the dimerization constants: KAH+/AH+ = 3 × 104, KAH+/A = 8 × 104, KA/A = 5 × 104 M–1.
Figure 18
Figure 18. Schematic (1D) potential energy surface (PES) showing many local minima accessible by 1 ns-long MD simulation and the necessity to properly explore the entire PES to reach the real valley via long MD simulation or biased techniques to force random exploration. (The noncovalent complex represented here is just a pictorial representation.)
Scheme 3
Scheme 3. Thermodynamics Cycle for the Formation of Copigmentation Complexes in Water (Comprising Complex Formation in the Gas Phase and the Transfer of the Process from the Gas Phase to Solvent)Figure 19
Figure 19. Molecular dynamics snapshot of the 3-OMe-cyanidin···quercetin noncovalent complex (the pigment 3-OMe-cyanidin in red; the copigment quercetin in yellow), showing the space between both partners depleted from water molecules.
Figure 20
Figure 20. π–π Stacking complexes between quercetin (yellow) and vitamin E (green) in lipid bilayer membranes.
Figure 21
Figure 21. Dispersion energy as a function of the interatomic distance for a set of selected atomic pairs.
Figure 22
Figure 22. Sketch of the two main DFT-based corrections discussed: the statically averaged interaction (top) and the dynamically interacting densities (bottom).
Figure 23
Figure 23. Five stable conformations of the 3-OMe-cyanidin··· quercetin noncovalent complex (A) side-view and top view showing the parallel-displaced arrangement; (B) side-view showing the intermolecular hydrogen bonding of conformers 2 and 5; (C) top view showing the supramolecular chirality in conformers 3 and 5; and (D) side-view showing the slight bending in conformers 1 and 5.
Figure 24
Figure 24. Potential energy curve along the z-axis defined perpendicular to the planar π-conjugated system, as obtained by three methods of calculation.
Scheme 4
Scheme 4. (A) Simplified MO diagram for CT and exciton formation of a DA and AA pair (H = HOMO, L = LUMO)aScheme aGrey arrows indicate lowest electronic transitions in a one-electron picture. (B) Resulting exciton states for rotated and laterally displaced AA pair (qualitative quantum-chemical corrected Kasha picture at π-stacking distance).
Figure 25
Figure 25. MO orbital diagram related to the maximum absorption wavelength HOMO and LUMO (left) and NTO analysis (right). Both schemes show the strong CT character of S0 → S1.
References
This article references 362 other publications.
- 1Yoshida, K.; Mori, M.; Kondo, T. Blue Flower Color Development by Anthocyanins: From Chemical Structure to Cell Physiology Nat. Prod. Rep. 2009, 26 (7) 857– 964 DOI: 10.1039/b800165kThere is no corresponding record for this reference.
- 2Boulton, R. The Copigmentation of Anthocyanins and Its Role in the Color of Red Wine: A Critical Review Am. J. Enol. Vitic. 2001, 52 (2) 67– 872The copigmentation of anthocyanins and its role in the color of red wine: A critical reviewBoulton, RogerAmerican Journal of Enology and Viticulture (2001), 52 (2), 67-87CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review with refs. Copigmentation is a soln. phenomenon in which pigments and other noncolored org. components form mol. assocns. or complexes. It generally results in an enhancement in the absorbance and in some cases, a shift in the wavelength of the max. absorbance of the pigment. Copigmentation has not previously been taken into account in traditional wine color measures, in the relationship between color and pigment anal., or in spectrophotometric assays for anthocyanin content. It is now apparent that copigmentation can account for between 30 and 50% of the color in young wines and that it is primarily influenced by the levels of several specific, noncolored phenolic components or cofactors. Copigmentation is of crit. importance in understanding the relationship between grape compn. and wine color, the variation in color and pigment concn. between wines, and in all reactions involving the anthocyanins during wine aging. This review focuses on the importance of the individual pigments and cofactors, the strength of their interactions, and their relative abundance in grapes and wines. A simple math. anal. of the soln. equil. is developed to explain the nonlinear deviation from Beer's law. When solved for typical wines, this function provides ests. of the apparent assocn. const., K, and the apparent molar extinction of the copigmented form, Ec, in natural mixts. These measures allow the fraction of the anthocyanins which is in the copigmented form to be estd. The significance of this phenomenon on pigment extn. and color retention during fermns., on the rate of subsequent pigment polymn., on the possible protection of anthocyanins from oxidn., and in the possible involvement on perceived mouth-feel and astringency of wines are suggested. Aspects of the copigmentation phenomenon that are poorly understood are identified and some research directions are suggested.
- 3Landrum, J. T. Carotenoids: Physical, Chemical, and Biological Functions and Properties; CRC Press: Boca Raton, 2010.There is no corresponding record for this reference.
- 4Haslam, E. Practical Polyphenolics; Cambridge University Press, 1998.There is no corresponding record for this reference.
- 5Lee, D. Nature’s Palette: The Science of Plant Color, Reprint ed.; University Of Chicago Press: Chicago, 2010.There is no corresponding record for this reference.
- 6Willstätter, R.; Everest, A. E. Untersuchungen Über Die Anthocyane. I. Über Den Farbstoff Der Kornblume Justus Liebigs Ann. Chem. 1913, 401 (2) 189– 232 DOI: 10.1002/jlac.19134010205There is no corresponding record for this reference.
- 7Asen, S.; Stewart, R. N.; Norris, K. H. Anthocyanin, Flavonol Copigments, and pH Responsible for Larkspur Flower Colour Phytochemistry 1975, 14 (12) 2677– 2682 DOI: 10.1016/0031-9422(75)85249-6There is no corresponding record for this reference.
- 8Castañeda-Ovando, A.; Pacheco-Hernández, M.; de, L.; Páez-Hernández, M. E.; Rodríguez, J. A.; Galán-Vidal, C. A. Chemical Studies of Anthocyanins: A Review Food Chem. 2009, 113 (4) 859– 871 DOI: 10.1016/j.foodchem.2008.09.0018Chemical studies of anthocyanins: A reviewCastaneda-Ovando, Araceli; Pacheco-Hernandez, Ma. de Lourdes; Paez-Hernandez, Ma. Elena; Rodriguez, Jose A.; Galan-Vidal, Carlos AndresFood Chemistry (2009), 113 (4), 859-871CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier B.V.)A review. Anthocyanins are natural colorants which have raised a growing interest due to their extensive range of colors, innocuous and beneficial health effects. Despite the great potential of application that anthocyanins represent for food, pharmaceutical and cosmetic industries, their use has been limited because of their relative instability and low extn. percentages. Currently, most investigations on anthocyanins are focused on solving these problems, as well as their purifn. and identification. In this paper, the most recent advances in the chem. investigation of the anthocyanins are summarized, emphasizing the effects of pH, co-pigmentation, metal ion complexation and antioxidant activity on their stability.
- 9He, F.; Liang, N.-N.; Mu, L.; Pan, Q.-H.; Wang, J.; Reeves, M. J.; Duan, C.-Q. Anthocyanins and Their Variation in Red Wines I. Monomeric Anthocyanins and Their Color Expression Molecules 2012, 17 (12) 1571– 1601 DOI: 10.3390/molecules17021571There is no corresponding record for this reference.
- 10Andersen, O. M.; Markham, K. R. Flavonoids: Chemistry, Biochemistry and Applications; CRC Press, 2005.There is no corresponding record for this reference.
- 11Fulcrand, H.; Dueñas, M.; Salas, E.; Cheynier, V. Phenolic Reactions during Winemaking and Aging Am. J. Enol. Vitic. 2006, 57 (3) 289– 29711Phenolic reactions during winemaking and agingFulcrand, Helene; Duenas, Montserrat; Salas, Erika; Cheynier, VeroniqueAmerican Journal of Enology and Viticulture (2006), 57 (3), 289-297CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review. The reactivity of polyphenols is due to the position of the hydroxyl groups on their arom. nuclei. Ortho-hydroxyl groups promote oxidn. while meta-hydroxyl groups induce electrophilic arom. substitution. Both hydroxylation patterns are encountered in flavonoid structures, on the B and A rings, resp. In addn. to oxidn. and electrophilic arom. substitution, flavonoids undergo nucleophilic addn. on the central C ring when it is pos. charged. Reactions of the A and C rings are pH-dependent. The A ring of flavonoids undergoes a polycondensation reaction mediated by an aldehyde. The products are anthocyanin and flavanol polymers and copolymers constituted of both. Flavanol polymers are not stable and rearrange into vinyl flavanols and xanthylium pigments. Vinyl flavanols can react with the pos. charged C ring of anthocyanins, yielding pyranoanthocyanins, which can also be formed from components that have a reactive double bond, such as carbonyl and ethylene bonds. The pos. charged C ring primarily undergoes direct reactions. Since the pos. charge on the C ring of anthocyanins and flavanols is pH-dependent, their dehydration and interflavan bond cleavage reactions are also pH-dependent. This leads to flavanol-anthocyanin (F-A+) adducts at lower pH values and anthocyanin-flavanol (A+-F) adducts above pH 3.8. Temp. seems to favor formation of the latter.
- 12Cheynier, V.; Dueñas-Paton, M.; Salas, E.; Maury, C.; Souquet, J.-M.; Sarni-Manchado, P.; Fulcrand, H. Structure and Properties of Wine Pigments and Tannins Am. J. Enol. Vitic. 2006, 57 (3) 298– 30512Structure and properties of wine pigments and tanninsCheynier, Veronique; Duenas-Paton, Montserrat; Salas, Erika; Maury, Chantal; Souquet, Jean-Marc; Sarni-Manchado, Pascale; Fulcrand, HeleneAmerican Journal of Enology and Viticulture (2006), 57 (3), 298-305CODEN: AJEVAC; ISSN:0002-9254. (American Society for Enology and Viticulture)A review. Grape phenolics are structurally diverse, from simple mols. to oligomers and polymers that are usually designated "tannins," referring to their ability to interact with proteins. Anthocyanin pigments and tannins are particularly important for red wine quality. Their extn. depends on their location in the berry and their soly. All phenolic compds. are unstable and undergo numerous enzymic and chem. reactions. Color and taste changes during red wine aging have been ascribed to anthocyanin-tannin reactions. The structures and properties of tannins and pigmented tannins from these reactions are often misunderstood. Current research on wine phenolic compn. is reviewed, with emphasis on the following issues: (1) reactions of tannins yield both larger polymers and smaller species; (2) anthocyanin reactions can generate colorless species as well as polymeric and small various pigments; (3) some polymeric pigments undergo sulfite bleaching while some low mol. wt. pigments do not; (4) polymers are both sol. and astringent, so the astringency loss during aging may involve cleavage rather than polymn.; and (5) sensory properties of anthocyanins and tannins are modulated by interactions with other wine components.
- 13de Freitas, V.; Mateus, N. Formation of Pyranoanthocyanins in Red Wines: A New and Diverse Class of Anthocyanin Derivatives Anal. Bioanal. Chem. 2011, 401 (5) 1463– 1473 DOI: 10.1007/s00216-010-4479-913Formation of pyranoanthocyanins in red wines: a new and diverse class of anthocyanin derivativesde Freitas Victor; Mateus NunoAnalytical and bioanalytical chemistry (2011), 401 (5), 1463-73 ISSN:.Pyranoanthocyanins constitute one of the most important classes of anthocyanin-derived pigments occurring naturally in red wine. Nonetheless, correct assignment of their structures and pathways of formation in red wine has been relatively recent--less than two decades. Study of these newly discovered pigments is progressively unfolding the chemical pathways that drive the evolution of red wine colour during ageing. The objective of this paper is to review current knowledge regarding the pathway of formation in red wine of a great variety of pyranoanthocyanin structures, namely carboxypyranoanthocyanins, methylpyranoanthocyanins, pyranoanthocyanin-flavanols, pyranoanthocyanin-phenols, portisins, oxovitisins, and pyranoanthocyanin dimers. The chromatic features of some of the compounds, for example their colour expression and acid-base equilibria in aqueous media, are also discussed.
- 14Mateus, N.; Silva, A. M. S.; Santos-Buelga, C.; Rivas-Gonzalo, J. C.; de Freitas, V. Identification of Anthocyanin-Flavanol Pigments in Red Wines by NMR and Mass Spectrometry J. Agric. Food Chem. 2002, 50 (7) 2110– 2116 DOI: 10.1021/jf0111561There is no corresponding record for this reference.
- 15Mateus, N.; Oliveira, J.; Haettich-Motta, M.; de Freitas, V. New Family of Bluish Pyranoanthocyanins J. Biomed. Biotechnol. 2004, 2004 (5) 299– 305 DOI: 10.1155/S1110724304404033There is no corresponding record for this reference.
- 16He, J.; Carvalho, A. R. F.; Mateus, N.; De Freitas, V. Spectral Features and Stability of Oligomeric Pyranoanthocyanin-Flavanol Pigments Isolated from Red Wines J. Agric. Food Chem. 2010, 58 (16) 9249– 9258 DOI: 10.1021/jf102085eThere is no corresponding record for this reference.
- 17He, J.; Oliveira, J.; Silva, A. M. S.; Mateus, N.; De Freitas, V. Oxovitisins: A New Class of Neutral Pyranone-Anthocyanin Derivatives in Red Wines J. Agric. Food Chem. 2010, 58 (15) 8814– 8819 DOI: 10.1021/jf101408qThere is no corresponding record for this reference.
- 18Chassaing, S.; Isorez, G.; Kueny-Stotz, M.; Brouillard, R. En Route to Color-Stable Pyranoflavylium Pigments—a Systematic Study of the Reaction between 5-Hydroxy-4-Methylflavylium Salts and Aldehydes Tetrahedron Lett. 2008, 49 (49) 6999– 7004 DOI: 10.1016/j.tetlet.2008.09.139There is no corresponding record for this reference.
- 19Oliveira, J.; Mateus, N.; Freitas, V. de. Synthesis of a New Bluish Pigment from the Reaction of a Methylpyranoanthocyanin with Sinapaldehyde Tetrahedron Lett. 2011, 52 (16) 1996– 2000 DOI: 10.1016/j.tetlet.2011.02.079There is no corresponding record for this reference.
- 20Brouillard, R.; Dubois, J.-E. Mechanism of the Structural Transformations of Anthocyanins in Acidic Media J. Am. Chem. Soc. 1977, 99 (5) 1359– 1364 DOI: 10.1021/ja00447a01220Mechanism of the structural transformations of anthocyanins in acidic mediaBrouillard, Raymond; Dubois, Jacques-EmileJournal of the American Chemical Society (1977), 99 (5), 1359-64CODEN: JACSAT; ISSN:0002-7863.In acidic aq. media (pH 1-6), there are three forms of malvin: the flavylium cation I (R = β-D-glucopyranosyl), the carbinol, and the quinonoidal base. Equilibrium between the two neutral forms occurs via the flavylium cation and the equil. const. [carbinol]/[base] is 1.6 (± 0.5) × 102 at 4° ;a relaxation technique was used to det. the rate consts for neutralization (endotherm) of the base, for hydration ofI via formation of a C-O bond and proton transfer,and for the reverse of the reactions.
- 21McClelland, R. A.; Gedge, S. Hydration of the Flavylium Ion J. Am. Chem. Soc. 1980, 102 (18) 5838– 5848 DOI: 10.1021/ja00538a02421Hydration of the flavylium ionMcClelland, Robert A.; Gedge, SherrinJournal of the American Chemical Society (1980), 102 (18), 5838-48CODEN: JACSAT; ISSN:0002-7863.Spectral and kinetic studies were carried out on the transformations undergone in aq. soln. by flavylium ions and I (R = H, Me, MeO). Seven species were identified under various pH conditions, i.e., the flavylium ion, 2 pseudobases, 4-hydroxy adducts and the cis-2-hydroxychalcones and their ionized forms. Equil. between the various species are described and rate and equil. consts. were detd.
- 22Pina, F.; Melo, M. J.; Laia, C. A. T.; Parola, A. J.; Lima, J. C. Chemistry and Applications of Flavylium Compounds: A Handful of Colours Chem. Soc. Rev. 2012, 41 (2) 869– 908 DOI: 10.1039/C1CS15126F22Chemistry and applications of flavylium compounds: a handful of coloursPina, Fernando; Melo, Maria J.; Laia, Cesar A. T.; Parola, A. Jorge; Lima, Joao C.Chemical Society Reviews (2012), 41 (2), 869-908CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Flavylium compds. are versatile mols. that comprise anthocyanins, the ubiquitous colorants used by Nature to confer color to most flowers and fruits. They have found a wide range of applications in human technol., from the millenary color paints described by the Roman architect Vitruvius, to their use as food additives, combining color and antioxidant effects, and even as light absorbers in solar cells aiming at a greener solar energy conversion. Their rich complexity derives in part from their ability to switch between a variety of species (flavylium cations, neutral quinoidal bases, hemiketals and chalcones, and neg. charged phenolates) by means of external stimuli, such as pH, temp. and light. This crit. review describes (i) the historical advancements in the understanding of the equil. of their chem. reaction networks; (ii) their thermodn. and kinetics; (iii) the mechanisms underlying their color development, such as co-pigmentation and host-guest interactions; (iv) the photophysics and photochem. that lead to photochromism; and (v) applications in solar cells, models for optical memories, photochromic soft materials such as ionic liqs. and gels, and their properties in solid state materials (274 refs.).
- 23Pina, F. Chemical Applications of Anthocyanins and Related Compounds. A Source of Bioinspiration J. Agric. Food Chem. 2014, 62 (29) 6885– 6897 DOI: 10.1021/jf404869m23Chemical Applications of Anthocyanins and Related Compounds. A Source of BioinspirationPina, FernandoJournal of Agricultural and Food Chemistry (2014), 62 (29), 6885-6897CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)A review. Independently of the natural or synthetic origin, flavylium derivs. follow the same network of chem. reactions. Actually, the flavylium cation is stable only at low pH values. Increasing the pH gives rise to the formation of several species: quinoidal bases, hemiketal, cis- and trans-chalcones, and their deprotonated forms. A deep knowledge of the thermodn. and kinetics of these species is an essential tool to practical applications of these compds., in particular, in the domain of food chem. In this work the network of chem. reactions involving flavylium derivs. is presented, and the resp. thermodn. and kinetics are discussed in detail, including the math. expressions and a step-by-step procedure to calc. all of the rate and equil. consts. of the system. Examples of systems possessing a high or low cis-trans isomerization barrier are shown. Recent practical applications of anthocyanins and related compds. illustrate the potential of the flavylium-based family of compds.
- 24Nave, F.; Petrov, V.; Pina, F.; Teixeira, N.; Mateus, N.; de Freitas, V. Thermodynamic and Kinetic Properties of a Red Wine Pigment: Catechin-(4,8)-Malvidin-3- O -Glucoside J. Phys. Chem. B 2010, 114 (42) 13487– 13496 DOI: 10.1021/jp104749fThere is no corresponding record for this reference.
- 25Mora-Soumille, N.; Al Bittar, S.; Rosa, M.; Dangles, O. Analogs of Anthocyanins with a 3′,4′-Dihydroxy Substitution: Synthesis and Investigation of Their Acid–base, Hydration, Metal Binding and Hydrogen-Donating Properties in Aqueous Solution Dyes Pigm. 2013, 96 (1) 7– 15 DOI: 10.1016/j.dyepig.2012.07.00625Analogs of anthocyanins with a 3',4'-dihydroxy substitution: Synthesis and investigation of their acid-base, hydration, metal binding and hydrogen-donating properties in aqueous solutionMora-Soumille, Nathalie; Al Bittar, Sheiraz; Rosa, Maxence; Dangles, OlivierDyes and Pigments (2013), 96 (1), 7-15CODEN: DYPIDX; ISSN:0143-7208. (Elsevier Ltd.)Glycosides of hydroxylated flavylium ions are proposed as pertinent analogs of anthocyanins, a major class of polyphenolic plant pigments. Anthocyanins with a 3',4'-dihydroxy substitution on the B-ring (catechol nucleus) are esp. important for their metal chelating and electron-donating (antioxidant) capacities. In this work, an efficient chem. synthesis of 3',4'-dihydroxy-7-O-β-D-glucopyranosyloxyflavylium chloride and its aglycon is reported. Then, the ability of the two pigments to undergo proton transfer (formation of colored quinonoid bases) and add water (formation of a colorless chalcone) is investigated: at equil. the colored quinonoid bases (kinetic products) are present in very minor concns. (<10% of the total pigment concn.) compared to the colorless chalcone (thermodn. product). The glucopyranosyloxyflavylium ion appears significantly less acidic than the aglycon. The thermodn. of the overall sequence of flavylium-chalcone conversion is not affected by the β-D-glucosyl moiety while the kinetics appears slower by a factor ca. 8. Although the glucopyranosyl-oxyflavylium ion and its aglycon display similar affinities for Al3+, the Al3+-glucoside complex is more stable than the Al3+-aglycon complex due to the higher sensitivity of the latter to water addn. and conversion into the corresponding chalcone. Finally, the glucopyranosyl-oxyflavylium ion and its aglycon are compared for their ability to reduce the 1,1-diphenyl-2-picrylhydrazyl radical in a mildly acidic water/MeOH (1:1) mixt. as a first evaluation of their antioxidant activity. Glycosidation at C7-OH results in a lower rate const. of first electron transfer to DPPH and a lower stoichiometry (total no. of 1,1-diphenyl-2-picrylhydrazyl radicals reduced per pigment mol.). Anthocyanins are difficult to ext. from plants in substantial amt. However, the analogs investigated in this work are of easy access by chem. synthesis and express the physico-chem. properties typical of anthocyanins. They can thus be regarded as valuable models for investigating the coloring, metal-binding and antioxidant properties of these important natural pigments.
- 26Yan, Q.; Zhang, L.; Zhang, X.; Liu, X.; Yuan, F.; Hou, Z.; Gao, Y. Stabilization of Grape Skin Anthocyanins by Copigmentation with Enzymatically Modified Isoquercitrin (EMIQ) as a Copigment Food Res. Int. 2013, 50 (2) 603– 609 DOI: 10.1016/j.foodres.2011.04.00726Stabilization of grape skin anthocyanins by copigmentation with enzymatically modified isoquercitrin (EMIQ) as a copigmentYan, Qiuli; Zhang, Linhan; Zhang, Xiaofei; Liu, Xuan; Yuan, Fang; Hou, Zhanqun; Gao, YanxiangFood Research International (2013), 50 (2), 603-609CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)The thermal and light stability of grape skin anthocyanins with enzymically modified isoquercitrin (EMIQ) as a copigment was investigated at different pH levels of 3, 4 and 5. The ratios of anthocyanins to EMIQ were 2:1, 1:1, and 1:2 (wt./wt.), resp., in the thermal expts. at 90 °C, and EMIQ concns. (0.25, 0.5, and 1%, wt./wt.) were evaluated resp. in the light expts. Results revealed that the degrdn. of anthocyanins copigmented with EMIQ followed first-order reaction kinetics. The half life of anthocyanins extended significantly with the increase of EMIQ concn. (p < 0.05), moreover, the color stability increased due to the addn. of EMIQ as the total color difference values ΔE* were smaller for the copigmented anthocyanins. The magnitude of the bathochromic (λmax) shifted to the longest wavelength absorption band with the increasing copigment concn. for all pH levels. Results demonstrated that EMIQ was an effective copigment to stabilize grape skin anthocyanins.
- 27Fanzone, M.; González-Manzano, S.; Pérez-Alonso, J.; Escribano-Bailón, M. T.; Jofré, V.; Assof, M.; Santos-Buelga, C. Evaluation of Dihydroquercetin-3-O-Glucoside from Malbec Grapes as Copigment of Malvidin-3-O-Glucoside Food Chem. 2015, 175, 166– 173 DOI: 10.1016/j.foodchem.2014.11.12327Evaluation of dihydroquercetin-3-O-glucoside from Malbec grapes as copigment of malvidin-3-O-glucosideFanzone, Martin; Gonzalez-Manzano, Susana; Perez-Alonso, Joaquin; Escribano-Bailon, Maria Teresa; Jofre, Viviana; Assof, Mariela; Santos-Buelga, CelestinoFood Chemistry (2015), 175 (), 166-173CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Malbec is a wine grape variety of great phenolic potential characterized for its high levels of anthocyanins and dihydroflavonols. To evaluate the possible implication of dihydroflavonols in the expression of red wine color through reactions of copigmentation or condensation, assays were carried out in wine model systems with different malvidin-3-O-glucoside:dihydroquercetin-3-O-glucoside molar ratios. The addn. of increasing levels of dihydroquercetin-3-O-glucoside to a const. malvidin-3-O-glucoside concn. resulted in a hyperchromic effect assocd. with a darkening of the anthocyanin solns., greater quantity of color and visual satn., perceptible to the human eye. Copigmentation and thermodn. measurements showed that dihydroquercetin-3-O-glucoside can act as an anthocyanin copigment, similar to other usual wine components like flavanols or phenolic acids, although apparently less efficient than flavonols. The high levels of dihydroflavonols existing in Malbec wines in relation to other non-anthocyanin phenolics should make this family of compds. particularly important to explain the color expression in Malbec young red wines.
- 28Baranac, J. M.; Petranovic, N. A.; Dimitric-Markovic, J. M. Spectrophotometric Study of Anthocyan Copigmentation Reactions. 2. Malvin and the Nonglycosidized Flavone Quercetin J. Agric. Food Chem. 1997, 45 (5) 1694– 1697 DOI: 10.1021/jf960611428Spectrophotometric Study of Anthocyan Copigmentation Reactions. 2. Malvin and the Nonglycosidized Flavone QuercetinBaranac, Jelisaveta M.; Petranovic, Nadezda A.; Dimitric-Markovic, Jasmina M.Journal of Agricultural and Food Chemistry (1997), 45 (5), 1694-1697CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Using UV-vis spectrophotometry we have established that a process of copigmentation takes place between an anthocyan mol., malvin chloride (malvidin 3,5-diglucoside), and a nonglycosidized pentahydroxyflavone, quercetin (3,5,7,3',4'-pentahydroxyflavone). The kinetic and thermodn. parameters, by which the process is characterized, were correlated to the structure, i.e., the nature and position of the substituents in the interacting mols.
- 29Dimitrić-Marković, J. M.; Petranović, N. A.; Baranac, J. M. A Spectrophotometric Study of the Copigmentation of Malvin with Caffeic and Ferulic Acids J. Agric. Food Chem. 2000, 48 (11) 5530– 5536 DOI: 10.1021/jf000038vThere is no corresponding record for this reference.
- 30Dimitrić-Marković, J. M.; Baranac, J. M.; Brdaric, T. P. Electronic and Infrared Vibrational Analysis of Cyanidin-Quercetin Copigment Complex Spectrochim. Acta, Part A 2005, 62, 673– 680 DOI: 10.1016/j.saa.2005.02.03630Electronic and infrared vibrational analysis of cyanidin-quercetin copigment complexDimitric Markovic, Jasmina M.; Baranac, Jelisaveta M.; Brdaric, Tanja P.Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy (2005), 62A (1-3), 673-680CODEN: SAMCAS; ISSN:1386-1425. (Elsevier B.V.)Copigment complex formation between cyanidin and quercetin, in aq. buffered solns., was studied by electronic absorption and IR vibrational spectroscopies. It was found that the assocn. of cyanidin with quercetin occurred at pH 3.0 and pH 5.0 including cyanidin flavylium ion and anhydrobase transformation forms, resp. The obtained copigmentation const. values of K = 2726.7 (pH 3.0) and K = 1093.1 (pH 5.0) indicated good assocn. ability of the investigated mols. IR spectra revealed the existence of hydrogen bonds in the copigment complexes structures. The anal. of the deconvoluted IR spectra indicated several types of hydrogen bonds, differently formed: the H-O ··· H bonds with the corresponding bands around 3500 cm-1 and bonds formed via H3O+, oxonium, ion of the mols. with the corresponding bands below 3000 cm-1.
- 31Lambert, S. G.; Asenstorfer, R. E.; Williamson, N. M.; Iland, P. G.; Jones, G. P. Copigmentation between Malvidin-3-Glucoside and Some Wine Constituents and Its Importance to Colour Expression in Red Wine Food Chem. 2011, 125 (1) 106– 115 DOI: 10.1016/j.foodchem.2010.08.04531Copigmentation between malvidin-3-glucoside and some wine constituents and its importance to colour expression in red wineLambert, Stephanie G.; Asenstorfer, Robert E.; Williamson, Natalie M.; Iland, Patrick G.; Jones, Graham P.Food Chemistry (2011), 125 (1), 106-115CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Thermodn. parameters for intermol. copigmentation interactions involving malvidin-3-glucoside were detd. by UV/visible spectroscopy at wine pH (pH 3.6). These included assocn. consts., enthalpy and entropy changes, which were measured for chlorogenic acid, caffeic acid, quercetin, quercetin-3-glucoside, (-)-epicatechin, (+)-catechin, procyanidin dimer and seed tannin. Quercetin produced the strongest copigment (KCP = 2900 ± 1300), while the addn. of glucose at position 3 (quercetin-3-glucoside) reduced its effect by almost 10-fold. Malvidin-3-glucoside self-assocn. (KD = 3300 ± 300 mol-1 l) was thermodynamically favored over intermol. interaction with any of the copigments tested. No color enhancement due to self-assocn. was obsd. for malvidin-3-glucoside-derived pigments that cannot enter hydration reactions. In addn., malvidin-3-(6-O-p-coumaryl)glucoside did not show color enhancement suggesting that the p-coumaryl group prevents self-assocn. The malvidin-3-glucoside CD spectrum was not affected by indicated changes in malvidin-3-glucoside concn. These observations demonstrate that self-assocn. of malvidin-3-glucoside is more important than copigmentation in young red wine.
- 32Malaj, N.; De Simone, B. C.; Quartarolo, A. D.; Russo, N. Spectrophotometric Study of the Copigmentation of Malvidin 3-O-Glucoside with P-Coumaric, Vanillic and Syringic Acids Food Chem. 2013, 141 (4) 3614– 3620 DOI: 10.1016/j.foodchem.2013.06.017There is no corresponding record for this reference.
- 33Willstätter, R.; Zollinger, E. H. XVI. Über Die Farbstoffe Der Weintraube Und Der Heidelbeere, II Justus Liebigs Ann. Chem. 1917, 412 (2) 195– 216 DOI: 10.1002/jlac.19174120207There is no corresponding record for this reference.
- 34Robinson, G. M.; Robinson, R. A Survey of Anthocyanins. I Biochem. J. 1931, 25 (5) 1687– 1705 DOI: 10.1042/bj0251687There is no corresponding record for this reference.
- 35Asen, S.; Stewart, R. N.; Norris, K. H.; Massie, D. R. A Stable Blue Non-Metallic Co-Pigment Complex of Delphanin and C-Glycosylflavones in Prof. Blaauw Iris Phytochemistry 1970, 9 (3) 619– 627 DOI: 10.1016/S0031-9422(00)85702-7There is no corresponding record for this reference.
- 36Asen, S.; Stewart, R. N.; Norris, K. H. Co-Pigmentation Effect of Quercetin Glycosides on Absorption Characteristics of Cyanidin Glycosides and Color of Red Wing Azalea Phytochemistry 1971, 10 (1) 171– 175 DOI: 10.1016/S0031-9422(00)90266-836Copigmentation effect of quercetin glycosides on absorption characteristics of cyanidin glycosides and color of red wing azaleaAsen, Sam; Stewart, Robert N.; Norris, Karl H.Phytochemistry (Elsevier) (1971), 10 (1), 171-5CODEN: PYTCAS; ISSN:0031-9422.A quercetin 5-methyl ether and 5 quercetin glycosides were isolated from flowers of Red Wing azalea but were found only in trace amts. in an orange sport of this cultivar. The anthocyanins (cyanidin glycosides) extd. from the orange and red flowers were identical, even though the absorbance spectra of the intact cells differed. The absorbance spectrum of the orange sport was simulated with a 10-3M aq. soln. of cyanidin 3,5-diglucoside at the pH of the tissue, 2.8. The absorbance spectrum of Red Wing was matched with cyanidin 3,5-diglucoside at the same concn. and pH, copigmented with the 3-rhamnoside or galactoside of quercetin.
- 37Asen, S.; Stewart, R. N.; Norris, K. H. Co-Pigmentation of Anthocyanins in Plant Tissues and Its Effect on Color Phytochemistry 1972, 11 (3) 1139– 1144 DOI: 10.1016/S0031-9422(00)88467-837Copigmentation of anthocyanins in plant tissues and its effect on colorAsen, S.; Stewart, R. N.; Norris, K. H.Phytochemistry (Elsevier) (1972), 11 (3), 1139-44CODEN: PYTCAS; ISSN:0031-9422.Glycosides of the 6 common anthocyanidins all formed copigment complexes with flavonoids and other compds. at pH 2-5. The formation of copigment complexes resulted in a bathochromic shift in the visible λmax. of the anthocyanins and a large increase in extinction at pH 3 and higher. These complexes apparently formed with both the red flavylium salts and the purple anhydro bases. The increase in extinction at pH 3-5 was attributed to the stabilizing effect copigmentation had on the anhydro bases. The degree of copigmentation was a function of the concn. of the anthocyanins and the molar ratio of copigments to anthocyanins. Copigmentation offers an explanation for the infinite color variations that occur in flowers in a pH range where anthocyanins alone are virtually colorless.
- 38Asen, S.; Norris, K. H.; Stewart, R. N. Copigmentation of Aurone and Flavone from Petals of Antirrhinum Majus Phytochemistry 1972, 11 (9) 2739– 2741 DOI: 10.1016/S0031-9422(00)86505-XThere is no corresponding record for this reference.
- 39Hoshino, T.; Matsumoto, U.; Goto, T. Self-Association of Some Anthocyanins in Neutral Aqueous Solution Phytochemistry 1981, 20 (8) 1971– 1976 DOI: 10.1016/0031-9422(81)84047-2There is no corresponding record for this reference.
- 40Hoshino, T.; Matsumoto, U.; Harada, N.; Goto, T. Chiral Exciton Coupled Stacking of Anthocyanins: Interpretation of the Origin of Anomalous CD Induced by Anthocyanin Association Tetrahedron Lett. 1981, 22 (37) 3621– 3624 DOI: 10.1016/S0040-4039(01)81976-6There is no corresponding record for this reference.
- 41Hoshino, T.; Matsumoto, U.; Goto, T. Evidences of the Self-Association of Anthocyanins I. Circular Dichroism of Cyanin Anhydrobase Tetrahedron Lett. 1980, 21 (18) 1751– 1754 DOI: 10.1016/S0040-4039(00)77827-0There is no corresponding record for this reference.
- 42Hoshino, T.; Matsumoto, U.; Goto, T.; Harada, N. Evidence for the Self-Association of Anthocyanins IV. PMR Spectroscopic Evidence for the Vertical Stacking of Anthocyanin Molecules Tetrahedron Lett. 1982, 23 (4) 433– 436 DOI: 10.1016/S0040-4039(00)86852-5There is no corresponding record for this reference.
- 43Hoshino, T.; Goto, T. Effects of pH and Concentration on the Self-Association of Malvin Quinonoidal Base: Electronic and Circular Dichroic Studies Tetrahedron Lett. 1990, 31 (11) 1593– 1596 DOI: 10.1016/0040-4039(90)80025-HThere is no corresponding record for this reference.
- 44Hoshino, T. Self-Association of Flavylium Cations of Anthocyanidin 3,5-Diglucosides Studied by Circular Dichroism and 1H NMR Phytochemistry 1992, 31 (2) 647– 653 DOI: 10.1016/0031-9422(92)90053-S44Anthocyanin self-aggregates. Part 7. Self-association of flavylium cations of anthocyanidin 3,5-diglucosides studied by circular dichroism and proton NMRHoshino, TsutomuPhytochemistry (1992), 31 (2), 647-53CODEN: PYTCAS; ISSN:0031-9422.Exciton coupling type CD was induced by increasing anthocyanin concn. in strongly acidic media. The higher the concn., the larger the CD magnitude. All anthocyanidin 3,5-diglucosides exhibited neg. signs in the first Cotton effect, which indicates that all the anthocyanins self-assoc. in a left-handed screw manner. Concn. dependence of proton chem. shifts were also obsd. with increase of anthocyanidin 3,5-diglucosides in a strongly acidic medium; upfield shifts of arom. rings by increasing pigment concns. demonstrated that anthocyanidin nuclei stack vertically. Analyses of proton chem. shifts based on the Dimicoli and Helene equation made possible the detn. of the self-assocn. of flavylium cations. The self-assocn. consts. for the most common natural anthocyanidin 3,5-diglucosides are given.
- 45Hoshino, T. An Approximate Estimate of Self-Association Constants and the Self-Stacking Conformation of Malvin Quinonoidal Bases Studied by 1H NMR Phytochemistry 1991, 30 (6) 2049– 2055 DOI: 10.1016/0031-9422(91)85065-8There is no corresponding record for this reference.
- 46Brouillard, R.; Mazza, G.; Saad, Z.; Albrecht-Gary, A. M.; Cheminat, A. The Co-Pigmentation Reaction of Anthocyanins: A Microprobe for the Structural Study of Aqueous Solutions J. Am. Chem. Soc. 1989, 111 (7) 2604– 2610 DOI: 10.1021/ja00189a03946The co-pigmentation reaction of anthocyanins: a microprobe for the structural study of aqueous solutionsBrouillard, R.; Mazza, G.; Saad, Z.; Albrecht-Gary, A. M.; Cheminat, A.Journal of the American Chemical Society (1989), 111 (7), 2604-10CODEN: JACSAT; ISSN:0002-7863.Visible absorption spectrometry shows that, in acidic aq. solns., chlorogenic acid (I) forms a loose 1:1 complex with malvin chloride (II). The mol. interaction taking place between these two chem. species is characteristic of the copigmentation reaction of anthocyanins. For the first time the mechanism assocd. with this reaction is established. The equation describing the copigment effect is also given. The copigmentation reaction is a very fast process which is extremely influenced by temp. Increasing the temp. or adding MeOH, HCONH2, or NaCl always reduces the copigment effect. In fact, the extent of copigmentation is strictly under the control of the unique mol. structure of liq. water. Finally, the copigmentation phenomenon, which is widespread in higher plants, constitutes a simple, inexpensive, and very sensitive, microprobe for the structural studies of aq. solns.
- 47Figueiredo, P.; Elhabiri, M.; Saito, N.; Brouillard, R. Anthocyanin Intramolecular Interactions. A New Mathematical Approach To Account for the Remarkable Colorant Properties of the Pigments Extracted from Matthiola Incana J. Am. Chem. Soc. 1996, 118 (20) 4788– 4793 DOI: 10.1021/ja9535064There is no corresponding record for this reference.
- 48Figueiredo, P.; Elhabiri, M.; Toki, K.; Saito, N.; Dangles, O.; Brouillard, R. New Aspects of Anthocyanin Complexation. Intramolecular Copigmentation as a Means for Colour Loss? Phytochemistry 1996, 41 (1) 301– 308 DOI: 10.1016/0031-9422(95)00530-7There is no corresponding record for this reference.
- 49Zhang, B.; Liu, R.; He, F.; Zhou, P.-P.; Duan, C.-Q. Copigmentation of Malvidin-3-O-Glucoside with Five Hydroxybenzoic Acids in Red Wine Model Solutions: Experimental and Theoretical Investigations Food Chem. 2015, 170, 226– 233 DOI: 10.1016/j.foodchem.2014.08.02649Copigmentation of malvidin-3-O-glucoside with five hydroxybenzoic acids in red wine model solutions: Experimental and theoretical investigationsZhang, Bo; Liu, Rui; He, Fei; Zhou, Pan-Pan; Duan, Chang-QingFood Chemistry (2015), 170 (), 226-233CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)In the present research, the copigmentations of malvidin-3-O-glucoside with five hydroxybenzoic cofactors (p-hydroxybenzoic acid, protocatechuic acid, gallic acid, vanillic acid, and syringic acid) were investigated. The influence of the concn. of these cofactors and the reaction temp. was examd. The equil. const. (K), stoichiometric ratio (n) and the thermodn. parameters (ΔG°, ΔH°, ΔS°) related to the copigmentation were also reported here. Theor. calcns. were performed to identify the relative arrangement between the pigment and cofactors in the copigmentation complexes. Besides, the comparison of the relative binding free energies (ΔΔGbinding) derived from the theor. calcns. and exptl. data were made, and the binding strength of these copigmentation complexes was discussed with the interaction energies (ΔE). AIM anal. was also used to explore the main driving forces contributing to the copigmentation. In the comparison of the five studied cofactors, syringic acid had a stronger copigmentation effect than the other four phenolic acids investigated.
- 50Teixeira, N.; Cruz, L.; Brás, N. F.; Mateus, N.; Ramos, M. J.; de Freitas, V. Structural Features of Copigmentation of Oenin with Different Polyphenol Copigments J. Agric. Food Chem. 2013, 61 (28) 6942– 6948 DOI: 10.1021/jf401174b50Structural Features of Copigmentation of Oenin with Different Polyphenol CopigmentsTeixeira, Natercia; Cruz, Luis; Bras, Natercia F.; Mateus, Nuno; Ramos, Maria Joao; de Freitas, VictorJournal of Agricultural and Food Chemistry (2013), 61 (28), 6942-6948CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The copigmentation binding consts. (K) for the interaction of different copigments with oenin (major red wine anthocyanin) were detd. All tests were performed in a 12% ethanol citrate buffer soln. (0.2 M) at pH 3.5, with an ionic strength adjusted to 0.5 M by the addn. of sodium chloride. Over the past years, several copigmentation studies were made and many copigments were tested, but none of them included prodelphinidin B3 or a dimeric-type adduct like oenin-(O)-catechin, probably due to the difficulty in obtaining them. The data yielded from this study allowed concluding that (a) the presence of a pyrogallol group in the B ring of the flavan-3-ol structure slightly increases the copigmentation potential and (b) within all copigments tested oenin-(O)-catechin was revealed to be the best. According to computational studies performed on epicatechin/oenin, epigallocatechin/oenin, procyanidin B3/oenin, and oenin-(O)-catechin/oenin complexes, the ΔGbinding energy of the oenin-(O)-catechin/oenin complex is the most neg. compared to the other copigmentation complexes, hence being more stable and thermodynamically favored. All structural data show that oenin-(O)-catechin and epigallocatechin are closer to the pigment mol., which is in accordance with these two copigments having the highest exptl. copigmentation binding consts. for oenin.
- 51Fossen, T.; Rayyan, S.; Holmberg, M. H.; Nimtz, M.; Andersen, Ø. M. Covalent Anthocyanin–flavone Dimer from Leaves of Oxalis Triangularis Phytochemistry 2007, 68 (5) 652– 662 DOI: 10.1016/j.phytochem.2006.10.030There is no corresponding record for this reference.
- 52Saito, N.; Nakamura, M.; Shinoda, K.; Murata, N.; Kanazawa, T.; Kato, K.; Toki, K.; Kasai, H.; Honda, T.; Tatsuzawa, F. Covalent Anthocyanin–flavonol Complexes from the Violet-Blue Flowers of Allium “Blue Perfume. Phytochemistry 2012, 80, 99– 108 DOI: 10.1016/j.phytochem.2012.04.011There is no corresponding record for this reference.
- 53Hayashi, K.; Abe, Y.; Mitsui, S. Blue Anthocyanin from the Flowers of Commelina, the Crystallisation and Some Properties Thereof Proc. Jpn. Acad. 1958, 34 (6) 373– 37853Anthocyanins. XXX. Blue anthocyanin from flowers of Commelina, the crystallization and some properties thereofHayashi, Kozo; Abe, Yukihide; Mitsui, SeijiProceedings of the Japan Academy (1958), 34 (), 373-8CODEN: PJACAW; ISSN:0021-4280.cf. C.A. 50, 16935g. Brilliant blue, prismatic crystals obtained after repeated fractionation and pptn. with water and EtOH were subjected to qual. chem. analysis and behavior. The substance contained an appreciable amt. of metallic elements of which Mg and K were inherent to the pigment mol. It is readily sol. in water but does not retain its color in org. solvents and is neither acid nor base. Paper electrophoresis indicated the existence of essential structural difference between the blue and red forms of anthocyanin. The org. components seemed to be delphinidin, glucose, and p-coumaric acid.
- 54Shibata, K.; Shibata, Y.; Kasiwagi, I. Studies on Anthocyanins: Color Variationin Anthocyanins J. Am. Chem. Soc. 1919, 41 (2) 208– 220 DOI: 10.1021/ja01459a008There is no corresponding record for this reference.
- 55Bloor, S. J. Blue Flower Colour Derived from Flavonol-Anthocyanin Co-Pigmentation in Ceanothus Papillosus Phytochemistry 1997, 45 (7) 1399– 1405 DOI: 10.1016/S0031-9422(97)00129-5There is no corresponding record for this reference.
- 56Tamura, H.; Kondo, T.; Goto, T. The Composition of Commelinin, a Highly Associated Metalloanthocaynin Present in the Blue Flower Petals of Commelina Communis Tetrahedron Lett. 1986, 27 (16) 1801– 1804 DOI: 10.1016/S0040-4039(00)84379-8There is no corresponding record for this reference.
- 57Kondo, T.; Ueda, M.; Isobe, M.; Goto, T. A New Molecular Mechanism of Blue Color Development with Protocyanin, a Supramolecular Pigment from Cornflower, Centaurea Cyanus Tetrahedron Lett. 1998, 39 (45) 8307– 8310 DOI: 10.1016/S0040-4039(98)01858-9There is no corresponding record for this reference.
- 58Kondo, T.; Oyama, K.; Yoshida, K. Chiral Molecular Recognition on Formation of a Metalloanthocyanin: A Supramolecular Metal Complex Pigment from Blue Flowers of Salvia Patens Angew. Chem., Int. Ed. 2001, 40 (5) 894– 897 DOI: 10.1002/1521-3773(20010302)40:5<894::AID-ANIE894>3.3.CO;2-#There is no corresponding record for this reference.
- 59Mori, M.; Kondo, T.; Yoshida, K. Cyanosalvianin, a Supramolecular Blue Metalloanthocyanin, from Petals of Salvia Uliginosa Phytochemistry 2008, 69 (18) 3151– 3158 DOI: 10.1016/j.phytochem.2008.03.015There is no corresponding record for this reference.
- 60Yoshida, K.; Toyama-Kato, Y.; Kameda, K.; Kondo, T. Sepal Color Variation of Hydrangea Macrophylla and Vacuolar pH Measured with a Proton-Selective Microelectrode Plant Cell Physiol. 2003, 44 (3) 262– 268 DOI: 10.1093/pcp/pcg033There is no corresponding record for this reference.
- 61Kondo, T.; Toyama-Kato, Y.; Yoshida, K. Essential Structure of Co-Pigment for Blue Sepal-Color Development of Hydrangea Tetrahedron Lett. 2005, 46 (39) 6645– 6649 DOI: 10.1016/j.tetlet.2005.07.146There is no corresponding record for this reference.
- 62Tanaka, M.; Fujimori, T.; Uchida, I.; Yamaguchi, S.; Takeda, K. A Malonylated Anthocyanin and Flavonols in Blue Meconopsis Flowers Phytochemistry 2001, 56 (4) 373– 376 DOI: 10.1016/S0031-9422(00)00357-5There is no corresponding record for this reference.
- 63Yoshida, K.; Kitahara, S.; Ito, D.; Kondo, T. Ferric Ions Involved in the Flower Color Development of the Himalayan Blue Poppy, Meconopsis Grandis Phytochemistry 2006, 67 (10) 992– 998 DOI: 10.1016/j.phytochem.2006.03.013There is no corresponding record for this reference.
- 64Markham, K. R.; Mitchell, K. A.; Boase, M. R. Malvidin-3-O-Glucoside-5-O-(6-Acetylglucoside) and Its Colour Manifestation in “Johnson”s Blue’ and Other “Blue” Geraniums Phytochemistry 1997, 45 (2) 417– 423 DOI: 10.1016/S0031-9422(96)00831-XThere is no corresponding record for this reference.
- 65Yabuya, T.; Nakamura, M.; Iwashina, T.; Yamaguchi, M.; Takehara, T. Anthocyanin-Flavone Copigmentation in Bluish Purple Flowers of Japanese Garden Iris (Iris Ensata Thunb.) Euphytica 1997, 98 (3) 163– 167 DOI: 10.1023/A:1003152813333There is no corresponding record for this reference.
- 66Bimpilas, A.; Tsimogiannis, D.; Balta-Brouma, K.; Lymperopoulou, T.; Oreopoulou, V. Evolution of Phenolic Compounds and Metal Content of Wine during Alcoholic Fermentation and Storage Food Chem. 2015, 178, 164– 171 DOI: 10.1016/j.foodchem.2015.01.09066Evolution of phenolic compounds and metal content of wine during alcoholic fermentation and storageBimpilas, Andreas; Tsimogiannis, Dimitrios; Balta-Brouma, Kalliopi; Lymperopoulou, Theopisti; Oreopoulou, VassilikiFood Chemistry (2015), 178 (), 164-171CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Changes in the principal phenolic compds. and metal content during the vinification process and storage under modified atm. (50% N2, 50% CO2) of Merlot and Syrah wines, from grapes cultivated in Greece, have been investigated. Comparing the variation of metals at maceration process, with the variation of monomeric anthocyanins and flavonols, an inverse relationship was noticed, that can be attributed to complexing reactions of polyphenols with particular trace elements. Cu decreased rapidly, whereas a similar behavior that could be expected for Fe and Mn was not confirmed. Differences in the profile of anthocyanins and flavonols in the fresh Merlot and Syrah wines are reported. During 1 yr of storage monomeric anthocyanins declined almost tenfold, probably due to polymn. reactions and copigmentation. Also, a decrease in flavonol glycosides and increase in the resp. aglycons was obsd., attributed to enzymic hydrolysis. The concn. of total phenols and all metals remained practically const.
- 67Hermosín Gutiérrez, I. Influence of Ethanol Content on the Extent of Copigmentation in a Cencibel Young Red Wine J. Agric. Food Chem. 2003, 51 (14) 4079– 4083 DOI: 10.1021/jf021029kThere is no corresponding record for this reference.
- 68Monagas, M.; Bartolomé, B. Anthocyanins and Anthocyanin-Derived Compounds. In Wine Chemistry and Biochemistry; Moreno-Arribas, M. V.; Polo, M. C., Eds.; Springer: New York, 2009; pp 439– 462.There is no corresponding record for this reference.
- 69Rustioni, L.; Bedgood, D. R.; Failla, O.; Prenzler, P. D.; Robards, K. Copigmentation and Anti-Copigmentation in Grape Extracts Studied by Spectrophotometry and Post-Column-Reaction HPLC Food Chem. 2012, 132 (4) 2194– 2201 DOI: 10.1016/j.foodchem.2011.12.058There is no corresponding record for this reference.
- 70De Rosso, M.; Tonidandel, L.; Larcher, R.; Nicolini, G.; Dalla Vedova, A.; De Marchi, F.; Gardiman, M.; Giust, M.; Flamini, R. Identification of New Flavonols in Hybrid Grapes by Combined Liquid Chromatography–mass Spectrometry Approaches Food Chem. 2014, 163, 244– 251 DOI: 10.1016/j.foodchem.2014.04.110There is no corresponding record for this reference.
- 71Kelebek, H.; Canbas, A.; Selli, S. HPLC-DAD–MS Analysis of Anthocyanins in Rose Wine Made From Cv. Öküzgözü Grapes, and Effect of Maceration Time on Anthocyanin Content Chromatographia 2007, 66 (3–4) 207– 212 DOI: 10.1365/s10337-007-0277-8There is no corresponding record for this reference.
- 72Gordillo, B.; Rodríguez-Pulido, F. J.; González-Miret, M. L.; Quijada-Morín, N.; Rivas-Gonzalo, J. C.; García-Estévez, I.; Heredia, F. J.; Escribano-Bailón, M. T. Application of Differential Colorimetry To Evaluate Anthocyanin–Flavonol–Flavanol Ternary Copigmentation Interactions in Model Solutions J. Agric. Food Chem. 2015, 63 (35) 7645– 7653 DOI: 10.1021/acs.jafc.5b00181There is no corresponding record for this reference.
- 73Somers, T. C.; Vérette, E. Phenolic Composition of Natural Wine Types. In Wine Analysis; Linskens, P. D. H.-F.; Jackson, P. D. J. F., Eds.; Modern Methods of Plant Analysis; Springer: Berlin, 1988; pp 219– 257.There is no corresponding record for this reference.
- 74González-Manzano, S.; Dueñas, M.; Rivas-Gonzalo, J. C.; Escribano-Bailón, M. T.; Santos-Buelga, C. Studies on the Copigmentation between Anthocyanins and Flavan-3-Ols and Their Influence in the Colour Expression of Red Wine Food Chem. 2009, 114 (2) 649– 656 DOI: 10.1016/j.foodchem.2008.10.00274Studies on the copigmentation between anthocyanins and flavan-3-ols and their influence in the colour expression of red wineGonzalez-Manzano, Susana; Duenas, Montserrat; Rivas-Gonzalo, Julian C.; Escribano-Bailon, M. Teresa; Santos-Buelga, CelestinoFood Chemistry (2009), 114 (2), 649-656CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier B.V.)With the aim of evaluating the importance of the copigmentation process between anthocyanins and flavanols on the color expression of red wine, assays were carried out in wine model systems with mixts. of compds. obtained from two Vitis vinifera grape varieties (Graciano and Tempranillo). Spectrophotometric and chromatic analyses were performed to evaluate the magnitude of the copigmentation and the modifications induced in the color of the solns. Measurement of the changes in the anthocyanin hydration const. (Kh) was also used to det. the strength of the copigmentation process. All the flavanols assayed induced significant changes in the color, perceptible to the human eye, of the wine-like anthocyanin solns. at concns. similar to those that can exist in red wines. The percentage contribution of the copigmentation with flavanols to the color of the anthocyanin solns. was found to range from 2% to 20%. The extent of this effect was related not only to the concn. of flavanols but also to the qual. compn. of the flavanol prepns., as influenced by the part of the grape (either skin or seed) and the variety considered. Divergences were found between the evaluation of the copigmentation based on chromatic parameters in the CIELAB color space and that based on the measurement at visible λ max, as the latter does not consider the integral color changes produced in the visible spectrum. The results obtained confirmed the importance of the qual. phenolic compn., detd. in the wine by the type of grape and winemaking practices, to the prodn. of an effective copigmentation process.
- 75Fernandes, A.; Brás, N. F.; Mateus, N.; Freitas, V. de. A Study of Anthocyanin Self-Association by NMR Spectroscopy New J. Chem. 2015, 39 (4) 2602– 2611 DOI: 10.1039/C4NJ02339KThere is no corresponding record for this reference.
- 76Heras-Roger, J.; Pomposo-Medina, M.; Díaz-Romero, C.; Darias-Martín, J. Copigmentation, Colour and Antioxidant Activity of Single-Cultivar Red Wines Eur. Food Res. Technol. 2014, 239 (1) 13– 19 DOI: 10.1007/s00217-014-2185-076Copigmentation, colour and antioxidant activity of single-cultivar red winesHeras-Roger, J.; Pomposo-Medina, M.; Diaz-Romero, C.; Darias-Martin, J.European Food Research and Technology (2014), 239 (1), 13-19CODEN: EFRTFO; ISSN:1438-2377. (Springer)A hundred and thirty-six single-cultivar red wines of different vintages were collected from several wineries in the Canary Islands in order to study the magnitude of the copigmentation phenomenon and the antioxidant activity. The contribution of free anthocyanins, copigmented anthocyanins and polymeric pigments to the color of wine, as well as the total phenols, the antioxidant activity (2,2-diphenyl-1-picrylhydrazyl, DPPH method) and the chromatic characteristics of the wines were detd. The influence of ageing time and the climatic conditions on these parameters was also studied. The wines made with Merlot, Ruby Cabernet and Syrah cultivars showed the highest parameters of color, and the largest contribution to the copigmented anthocyanins was from the Ruby Cabernet, Listan negro and Syrah cultivars. The copigmented anthocyanins and the free anthocyanins decrease with the age of the wine, and the antioxidant activity of the samples appears to be related to the total phenol content. An influence of the climatic conditions on color parameters has been found. The correlation study between parameters suggests that the parameters b* and L* could be used as suitable indicators of evolution or oxidn. stage of red wines.
- 77García-Marino, M.; Escudero-Gilete, M. L.; Heredia, F. J.; Escribano-Bailón, M. T.; Rivas-Gonzalo, J. C. Color-Copigmentation Study by Tristimulus Colorimetry (CIELAB) in Red Wines Obtained from Tempranillo and Graciano Varieties Food Res. Int. 2013, 51 (1) 123– 131 DOI: 10.1016/j.foodres.2012.11.03577Color-copigmentation study by tristimulus colorimetry (CIELAB) in red wines obtained from Tempranillo and Graciano varietiesGarcia-Marino, Matilde; Escudero-Gilete, Maria Luisa; Heredia, Francisco Jose; Escribano-Bailon, Maria Teresa; Rivas-Gonzalo, Julian CarlosFood Research International (2013), 51 (1), 123-131CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)A study of the changes of copigmentation phenomenon in wines elaborated from different varieties has been undertaken. Colorimetric measurement of Tempranillo (T) and Graciano (G) monovarietal wines, and two 80:20 blend wines: M, (grape blending T and G, co-maceration) and W (wine blending T and G, co-vinification) was performed by spectrophotometry. Significant differences (p < 0.05) were found among the color of the wines. The Graciano cv. afforded somewhat darker and more colorful wines than the other wines. The color difference values, ΔE*ab suggested that co-vinification (W) led to wines being more similar to T than the co-maceration (M). The ΔE*ab[w-c] between untreated wines - whole wines, w - and the wines dild. to eliminate copigmentation - cor. wines, c - was 14.2 CIELAB units in the initial stages of winemaking and 6.7 in the final stages. M had a greater proportion of color due to copigmentation than the monovarietal wines. Evaluation of this parameter confirms the importance of copigmentation process into wine color during the early stages of the vinification. Also, through the full spectrum, quant. data obtained allow a visual interpretation of the changes involved. In addn., with the aging in bottle, M wines had more stable color and more different color than W wines.
- 78Schwarz, M.; Picazo-Bacete, J. J.; Winterhalter, P.; Hermosín-Gutiérrez, I. Effect of Copigments and Grape Cultivar on the Color of Red Wines Fermented after the Addition of Copigments J. Agric. Food Chem. 2005, 53 (21) 8372– 8381 DOI: 10.1021/jf051005oThere is no corresponding record for this reference.
- 79Aleixandre-Tudó, J. L.; Álvarez, I.; Lizama, V.; García, M. J.; Aleixandre, J. L.; Du Toit, W. J. Impact of Caffeic Acid Addition on Phenolic Composition of Tempranillo Wines from Different Winemaking Techniques J. Agric. Food Chem. 2013, 61 (49) 11900– 11912 DOI: 10.1021/jf402713dThere is no corresponding record for this reference.
- 80Gris, E. F.; Ferreira, E. A.; Falcão, L. D.; Bordignon-Luiz, M. T. Influence of Ferulic Acid on Stability of Anthocyanins from Cabernet Sauvignon Grapes in a Model System and a Yogurt System Int. J. Food Sci. Technol. 2007, 42 (8) 992– 998 DOI: 10.1111/j.1365-2621.2006.01335.x80Influence of ferulic acid on stability of anthocyanins from Cabernet Sauvignon grapes in a model system and a yogurt systemGris, Eliana Fortes; Ferreira, Eduardo Antonio; Falcao, Leila Denise; Bordignon-Luiz, Marilde TerezinhaInternational Journal of Food Science and Technology (2007), 42 (8), 992-998CODEN: IJFTEZ; ISSN:0950-5423. (Blackwell Publishing Ltd.)The influence of different factors on the stability of the anthocyanin crude ext. from Cabernet Sauvignon grape skins was investigated. In a model system, the factors evaluated were as follows: temp. 4 ± 1 °C and 29 ± 3 °C, presence and absence of light, pH 3.0 and 4.0 and presence of ferulic acid. The influence of the addn. of ferulic acid to anthocyanins was investigated in a yogurt system stored at 4 ± 1 °C. The results obtained for anthocyanin degrdn. velocity const. and for the half-life time of anthocyanins in a model system and in a yogurt system showed that ferulic acid significantly increased the stability of the anthocyanins crude ext.
- 81Talcott, S. T.; Peele, J. E.; Brenes, C. H. Red Clover Isoflavonoids as Anthocyanin Color Enhancing Agents in Muscadine Wine and Juice Food Res. Int. 2005, 38 (10) 1205– 1212 DOI: 10.1016/j.foodres.2005.05.00481Red clover isoflavonoids as anthocyanin color enhancing agents in muscadine wine and juiceTalcott, Stephen T.; Peele, Janelle E.; Brenes, Carmen H.Food Research International (2005), 38 (10), 1205-1212CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)Isoflavonoid exts. from red clover (Trifolium pratense) leaves were found to enhance overall color and stability of anthocyanin 3,5-diglucosides present in muscadine grape (Vitis rotundifolia) juice and wine through intermol. copigmentation reactions. Predominant isoflavonoids present in red clover included formononetin, biochanin A, and prunetin and were the major polyphenolics identified to influence anthocyanin color and stability. Since red clover isoflavonoids have poor water soly. characteristics, this allowed for removal of extraneous non-isoflavonoid compds. using hot water and subsequent extn. with ethanol. Isoflavonoid soly. was evaluated as a function of ethanol concn. with recoveries up to 57% found in 20% solns. Changes in max. absorbance, total sol. phenolics, isoflavonoids, and anthocyanins were evaluated in muscadine juice and wine following the addn. of isoflavonoid exts. with max. color enhancement found at an anthocyanin to cofactor ratio of 1:8, after which their soly. was prohibitive. Addnl., dried leaves and ethanolic exts. of red clover were added prior to and following fermn. of muscadine wine (11% ethanol) stimulating the natural copigmentation that takes place during red wine fermn. and aging processes. Once fermn. was complete, finished wines were evaluated over a 9-wk storage period at 20 and 37°. Despite low levels of isoflavonoids present, color improvement and anthocyanin stability was obsd. in the wines during storage. Little information is available on copigmentation reactions occurring in actual food systems, yet red clover isoflavonoids proved to be novel and effective color enhancing compds. when used in low concns. in young muscadine wines.
- 82Liu, S.; Fu, Y.; Nian, S. Buffering Colour Fluctuation of Purple Sweet Potato Anthocyanins to Acidity Variation by Surfactants Food Chem. 2014, 162, 16– 21 DOI: 10.1016/j.foodchem.2014.04.02982Buffering color fluctuation of purple sweet potato anthocyanins to acidity variation by surfactantsLiu, Songbai; Fu, Yuanqing; Nian, SiFood Chemistry (2014), 162 (), 16-21CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)Anthocyanins are intriguing natural pigments with beneficial bioactivities and their color is extremely susceptible to acidity variation. Minimization of color fluctuation is essential to maintain quality consistency in food industry. A new strategy employing surfactants to mimic encapsulation was attempted with typical anionic, cationic and nonionic surfactants and proved effective although the traditional copigmentation method was inactive. The exceptional color fluctuation buffering effect of anionic surfactants esp. SDS was revealed and then carefully analyzed by colorimetric and spectroscopic methods. The outstanding activity of SDS presumably resulted from effective shielding of anthocyanins from external acidity through strong interaction with the pos. charged flavylium cations owing to its anionic nature. These results suggest SDS is a valuable additive for buffering color fluctuation of anthocyanins. The strategy of surfactant will be useful for buffering color fluctuation of natural colorants.
- 83Kovac, V.; Alonso, E.; Bourzeix, M.; Revilla, E. Effect of Several Enological Practices on the Content of Catechins and Proanthocyanidins of Red Wines J. Agric. Food Chem. 1992, 40 (10) 1953– 1957 DOI: 10.1021/jf00022a045There is no corresponding record for this reference.
- 84Canals, R.; del Carmen Llaudy, M.; Canals, J. M.; Zamora, F. Influence of the Elimination and Addition of Seeds on the Colour, Phenolic Composition and Astringency of Red Wine Eur. Food Res. Technol. 2008, 226 (5) 1183– 1190 DOI: 10.1007/s00217-007-0650-884Influence of the elimination and addition of seeds on the colour, phenolic composition and astringency of red wineCanals, Roser; del Carmen Llaudy, Maria; Canals, Joan Miquel; Zamora, FernandoEuropean Food Research and Technology (2008), 226 (5), 1183-1190CODEN: EFRTFO; ISSN:1438-2377. (Springer GmbH)Several winemaking techniques were developed to eliminate seeds to prevent the release of high amts. of very astringent proanthocyanidins, esp. when the grapes are unripe. However, there is no scientific information on the effects of this practice. The aim of this paper is to study how the elimination and addn. of seeds influence the color, phenolic compn. and astringency of red wine. The elimination of around 80% of the seeds led to a significant decrease of color intensity and anthocyanin concn. The addn. of seeds originated wines with a greater concn. of total anthocyanin, but did not significantly affect the wine color. These wines also presented significantly higher levels of proanthocyanidins, a greater proportion of epicatechin-3-gallate, a lower mean d.p., and, in particular, a drastic increase in astringency. Wines obtained with the elimination of seeds, on the other hand, had exactly the opposite characteristics.
- 85Escribano-Bailon, M. T.; Santos-Buelga, C. Anthocyanin Copigmentation - Evaluation, Mechanisms and Implications for the Colour of Red Wines Curr. Org. Chem. 2012, 16 (6) 715– 723 DOI: 10.2174/13852721279995797785Anthocyanin copigmentation - evaluation, mechanisms and implications for the colour of red winesEscribano-Bailon, Maite T.; Santos-Buelga, CelestinoCurrent Organic Chemistry (2012), 16 (6), 715-723CODEN: CORCFE; ISSN:1385-2728. (Bentham Science Publishers Ltd.)A review. Copigmentation is the main color-stabilizing mechanism in plants and in food products of vegetable origin. It is a spontaneous and exothermic process that consists of the stacking of an org. mol., called copigment, on the planar polarizable moieties of the anthocyanin colored forms. Although this phenomenon has long been described, there are some aspects that are still not well understood or controversial like the nature of the interaction pigment to copigment, the way to quantify the extent of the process, its effect on other anthocyanin properties like astringency or reactivity. In this article a review of the most significant advances achieved in the last years in the field of intramol. and intermol. copigmentation is presented. Also, the most recent findings regarding wine copigments and their effects on the color of red wines are revised.
- 86Quijada-Morín, N.; Dangles, O.; Rivas-Gonzalo, J. C.; Escribano-Bailón, M. T. Physico-Chemical and Chromatic Characterization of Malvidin 3-Glucoside-Vinylcatechol and Malvidin 3-Glucoside-Vinylguaiacol Wine Pigments J. Agric. Food Chem. 2010, 58 (17) 9744– 9752 DOI: 10.1021/jf102238vThere is no corresponding record for this reference.
- 87Oliveira, J.; Mateus, N.; Silva, A. M. S.; de Freitas, V. Equilibrium Forms of Vitisin B Pigments in an Aqueous System Studied by NMR and Visible Spectroscopy J. Phys. Chem. B 2009, 113 (32) 11352– 11358 DOI: 10.1021/jp904776kThere is no corresponding record for this reference.
- 88Brouillard, R.; Dangles, O. Anthocyanin Molecular Interactions: The First Step in the Formation of New Pigments during Wine Aging? Food Chem. 1994, 51 (4) 365– 371 DOI: 10.1016/0308-8146(94)90187-2There is no corresponding record for this reference.
- 89Wrolstad, R. E.; Erlandson, J. A. Effect of Metal Ions on the Color of Strawberry Puree J. Food Sci. 1973, 38 (3) 460– 463 DOI: 10.1111/j.1365-2621.1973.tb01454.xThere is no corresponding record for this reference.
- 90Starr, M. S.; Francis, F. J. Effect of Metallic Ions on Color and Pigment Content of Cranberry Juice Cocktail J. Food Sci. 1973, 38 (6) 1043– 1046 DOI: 10.1111/j.1365-2621.1973.tb02144.xThere is no corresponding record for this reference.
- 91Kallio, H.; Pallasaho, S.; Kärppä, J.; Linko, R. R. Comparison of the Half-Lives of the Anthocyanins in the Juice of Crowberry, Empetrum Nigrum J. Food Sci. 1986, 51 (2) 408– 410 DOI: 10.1111/j.1365-2621.1986.tb11142.xThere is no corresponding record for this reference.
- 92Osawa, Y. Chapter 2 - Copigmentation of Anthocyanins. In Anthocyanins As Food Colors; Markakis, P., Ed.; Academic Press, 1982; pp 41– 68.There is no corresponding record for this reference.
- 93Rein, M. J.; Heinonen, M. Stability and Enhancement of Berry Juice Color J. Agric. Food Chem. 2004, 52 (10) 3106– 3114 DOI: 10.1021/jf035507iThere is no corresponding record for this reference.
- 94Maccarone, E.; Maccarrone, A.; Rapisarda, P. Stabilization of Anthocyanins of Blood Orange Fruit Juice J. Food Sci. 1985, 50 (4) 901– 904 DOI: 10.1111/j.1365-2621.1985.tb12976.xThere is no corresponding record for this reference.
- 95Wilska-Jeszka, J.; Korzuchowska, A. Anthocyanins and Chlorogenic Acid Copigmentation - Influence on the Colour of Strawberry and Chokeberry Juices Z. Lebensm.-Unters. Forsch. 1996, 203 (1) 38– 42 DOI: 10.1007/BF0126776795Anthocyanins and chlorogenic acid copigmentation. Influence on the color of strawberry and chokeberry juicesWilska-Jeszka, Jadwiga; Korzuchowska, AnnaZeitschrift fuer Lebensmittel-Untersuchung und -Forschung (1996), 203 (1), 38-42CODEN: ZLUFAR; ISSN:0044-3026. (Springer)The effect of copigmentation on chlorogenic acid with anthocyanins in strawberry and chokeberry juices was investigated. Chlorogenic acid, at concns. greater than that of anthocyanins, enhanced the color intensity of these juices. The max. copigmentation effect in both juices was obsd. at pH 3.4. In the investigated range of the copigment/pigment ratio, i.e. 1:1 to 50:1, absorbance increased (ΔA) linearly with copigment content, ΔA/g chlorogenic acid was greater in chokeberry than in strawberry juices. In solns. of purified pigments of these fruits, smaller copigmentation effects were obsd. than in juices under the same conditions, which indicated the participation of natural copigments present in fruits in the copigmentation process.
- 96Hernández-Herrero, J. A.; Frutos, M. J. Influence of Rutin and Ascorbic Acid in Colour, Plum Anthocyanins and Antioxidant Capacity Stability in Model Juices Food Chem. 2015, 173, 495– 500 DOI: 10.1016/j.foodchem.2014.10.059There is no corresponding record for this reference.
- 97Sari, P.; Wijaya, C. H.; Sajuthi, D.; Supratman, U. Colour Properties, Stability, and Free Radical Scavenging Activity of Jambolan (Syzygium Cumini) Fruit Anthocyanins in a Beverage Model System: Natural and Copigmented Anthocyanins Food Chem. 2012, 132 (4) 1908– 1914 DOI: 10.1016/j.foodchem.2011.12.025There is no corresponding record for this reference.
- 98Fischer, U. A.; Carle, R.; Kammerer, D. R. Thermal Stability of Anthocyanins and Colourless Phenolics in Pomegranate (Punica Granatum L.) Juices and Model Solutions Food Chem. 2013, 138 (2–3) 1800– 1809 DOI: 10.1016/j.foodchem.2012.10.072There is no corresponding record for this reference.
- 99Pan, Y.-Z.; Guan, Y.; Wei, Z.-F.; Peng, X.; Li, T.-T.; Qi, X.-L.; Zu, Y.-G.; Fu, Y.-J. Flavonoid C-Glycosides from Pigeon Pea Leaves as Color and Anthocyanin Stabilizing Agent in Blueberry Juice Ind. Crops Prod. 2014, 58, 142– 147 DOI: 10.1016/j.indcrop.2014.04.02999Flavonoid C-glycosides from pigeon pea leaves as color and anthocyanin stabilizing agent in blueberry juicePan, You-Zhi; Guan, Yue; Wei, Zuo-Fu; Peng, Xiao; Li, Ting-Ting; Qi, Xiao-Lin; Zu, Yuan-Gang; Fu, Yu-JieIndustrial Crops and Products (2014), 58 (), 142-147CODEN: ICRDEW; ISSN:0926-6690. (Elsevier B.V.)The influences of flavonoid C-glycoside exts. from pigeon pea leaves (FCGE) and its main components vitexin and orientin on the color and anthocyanins stability of blueberry juice were investigated in the thermal expts. The color of juice was enhanced by FCGE and its main components in a molar ratio of 1:1 (anthocyanins/copigment). The satd. color of juice with FCGE and its main components could hold for a significantly longer time. Addnl., copigment FCGE, vitexin and orientin significantly enhanced the stability of anthocyanins. Half-life of anthocyanins in juice samples with FCGE (1:1), orientin and vitexin increased 87%, 79%, and 62%, resp. These results indicated FCGE showed dramatic effect on the color and anthocyanins stability of juice. Furthermore, juice samples with copigments possessed higher total phenolics content and higher DPPH radical-scavenging activity. Therefore, natural, easily available, inexpensive FCGE has potential as color enhancer and anthocyanins stabilizer in food industry.
- 100Goto, T.; Kondo, T. Struktur Und Molekulare Stapelung von Anthocyanen—Variation Der Blütenfarben Angew. Chem. 1991, 103 (1) 17– 33 DOI: 10.1002/ange.19911030105There is no corresponding record for this reference.
- 101Liao, H.; Cai, Y.; Haslam, E. Polyphenol Interactions. Anthocyanins: Co-Pigmentation and Colour Changes in Red Wines J. Sci. Food Agric. 1992, 59 (3) 299– 305 DOI: 10.1002/jsfa.2740590305101Polyphenol interactions. Part 6. Anthocyanins: co-pigmentation and color changes in red winesLiao, Hua; Cai, Ya; Haslam, EdwinJournal of the Science of Food and Agriculture (1992), 59 (3), 299-305CODEN: JSFAAE; ISSN:0022-5142.Anthocyanin co-pigmentation is reviewed and its relevance to the color and appearance of young red wines is outlined. Reactions between flavan-3-ols, such as are present in red wines, and the anthocyanin malvin to give yellow-orange pigments (absorbance max. ∼440 nm) are reported. Possible structures for these new pigments are discussed as is their probable contribution to the changing hues of red wines as they age.
- 102Brouillard, R.; Wigand, M.-C.; Dangles, O.; Cheminat, A. pH and Solvent Effects on the Copigmentation Reaction of Malvin with Polyphenols, Purine and Pyrimidine Derivatives J. Chem. Soc., Perkin Trans. 2 1991, 8) 1235– 1241 DOI: 10.1039/p29910001235102The pH and solvent effects on the copigmentation reaction of malvin with polyphenols, purine and pyrimidine derivativesBrouillard, Raymond; Wigand, Marie Claude; Dangles, Olivier; Cheminat, AnnieJournal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1991), (8), 1235-41CODEN: JCPKBH; ISSN:0300-9580.The influence of pH on the copigmentation reaction of malvin has been investigated from an exptl. and theor. viewpoint. The general equation for the copigment effect, when monitored by visible absorption spectrometry, is derived and is in good agreement with results obtained in the case of three different copigments, namely chlorogenic acid, caffeine and adenosine. In particular, it is demonstrated that assocn. of malvin with the copigment occurs for all colored malvin species and the corresponding stability consts. are given. Some tannins and purine or pyrimidine derivs. have also been tested for their ability to act as copigments; some assoc. quite strongly with malvin. Water exerted the greatest effect on the extent of the copigmentation phenomenon. Such a result seems to indicate that the strength of the copigment effect parallels the cohesion of the H-bonded tetrahedral network of water mols.
- 103Baranac, J. M.; Petranovic, N. A.; Dimitric-Markovic, J. M. Spectrophotometric Study of Anthocyan Copigmentation Reactions J. Agric. Food Chem. 1996, 44 (5) 1333– 1336 DOI: 10.1021/jf950420lThere is no corresponding record for this reference.
- 104Oszmiański, J.; Bakowska, A.; Piacente, S. Thermodynamic Characteristics of Copigmentation Reaction of Acylated Anthocyanin Isolated from Blue Flowers ofScutellaria Baicalensis Georgi with Copigments J. Sci. Food Agric. 2004, 84 (12) 1500– 1506 DOI: 10.1002/jsfa.1815104Thermodynamic characteristics of copigmentation reaction of acylated anthocyanin isolated from blue flowers of Scutellaria baicalensis Georgi with copigmentsOszmianski, Jan; Bakowska, Anna; Piacente, SoniaJournal of the Science of Food and Agriculture (2004), 84 (12), 1500-1506CODEN: JSFAAE; ISSN:0022-5142. (John Wiley & Sons Ltd.)Anthocyanin exts. are increasingly used as food colorants. So far, anthocyanins have not been broadly used in foods and beverages, since they are not as stable as synthetic dyes. Copigmentation between anthocyanins and copigments is the main color-stabilizing mechanism. The process of copigmentation between isolated acylated anthocyanin and rutin, QSA or baicalin has been obsd. using UV-vis spectrophotometry. The thermodn. parameters were correlated to the structure and position of the substituents in the interacting mols. The acylated anthocyanin was isolated from cultivars of Scutellaria baicalensis Georgi flowers and purified by column chromatog. by our own method and has been identified by 1H-/13C-NMR spectroscopy and electrospray mass spectrometry as delphinidin-3-O-(6-O-malonyl)-β-D-glucopyranosyl-5-O-β-D-glucopyranoside.
- 105Ferreira da Silva, P.; Lima, J. C.; Freitas, A. A.; Shimizu, K.; Maçanita, A. L.; Quina, F. H. Charge-Transfer Complexation as a General Phenomenon in the Copigmentation of Anthocyanins J. Phys. Chem. A 2005, 109 (32) 7329– 7338 DOI: 10.1021/jp052106s105Charge-Transfer Complexation as a General Phenomenon in the Copigmentation of AnthocyaninsFerreira da Silva, Palmira; Lima, Joao C.; Freitas, Adilson A.; Shimizu, Karina; Macanita, Antonio L.; Quina, Frank H.Journal of Physical Chemistry A (2005), 109 (32), 7329-7338CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Color intensification of anthocyanin solns. in the presence of natural polyphenols (copigmentation) is re-interpreted in terms of charge transfer from the copigment to the anthocyanin. Flavylium cations are shown to be excellent electron acceptors (Ered ≈ -0.3 V vs SCE). It is also demonstrated, for a large series of anthocyanin-copigment pairs, that the std. Gibbs free energy of complex formation decreases linearly with EAAnthoc-IPCop, the difference between the electron affinity of the anthocyanin, EAAnthoc, and the ionization potential of the copigment, IPCop. Based on this correlation, copigmentation strengths of potential candidates for copigments can be predicted.
- 106Galland, S.; Mora, N.; Abert-Vian, M.; Rakotomanomana, N.; Dangles, O. Chemical Synthesis of Hydroxycinnamic Acid Glucosides and Evaluation of Their Ability To Stabilize Natural Colors via Anthocyanin Copigmentation J. Agric. Food Chem. 2007, 55 (18) 7573– 7579 DOI: 10.1021/jf071205v106Chemical synthesis of hydroxycinnamic acid glucosides and evaluation of their ability to stabilize natural colors via anthocyanin copigmentationGalland, Stephanie; Mora, Nathalie; Abert-Vian, Maryline; Rakotomanomana, Njara; Dangles, OlivierJournal of Agricultural and Food Chemistry (2007), 55 (18), 7573-7579CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)This work describes the chem. synthesis of O-aryl-β-D-glucosides and 1-O-β-D-glucosyl esters of hydroxycinnamic acids. In particular, O-aryl-β-D-glucosides were efficiently prepd. via a simple diastereoselective glycosylation procedure using phase transfer conditions. Despite the lability of its ester linkage, 1-O-β-D-caffeoylglucose could also be obtained using a Lewis acid catalyzed glycosylation step and a set of protective groups that can be removed under neutral conditions. Hydroxycinnamic acid O-aryl-β-D-glucosides were then quant. investigated for their affinity for the naturally occurring anthocyanin malvin (pigment). Formation of the π-stacking mol. complexes (copigmentation) was characterized in terms of binding consts. and enthalpy and entropy changes. The glucosyl moiety did not significantly alter these thermodn. parameters, in line with a binding process solely involving the polyphenolic nuclei.
- 107Alluis, B.; Dangles, O. Quercetin (=2-(3,4-Dihydroxyphenyl)-3,5,7-Trihydroxy-4H-1-Benzopyran-4-One) Glycosides and Sulfates: Chemical Synthesis, Complexation, and Antioxidant Properties Helv. Chim. Acta 2001, 84 (5) 1133– 1156 DOI: 10.1002/1522-2675(20010516)84:5<1133::AID-HLCA1133>3.3.CO;2-QThere is no corresponding record for this reference.
- 108Nave, F.; Brás, N. F.; Cruz, L.; Teixeira, N.; Mateus, N.; Ramos, M. J.; Di Meo, F.; Trouillas, P.; Dangles, O.; De Freitas, V. Influence of a Flavan-3-Ol Substituent on the Affinity of Anthocyanins (Pigments) toward Vinylcatechin Dimers and Proanthocyanidins (Copigments) J. Phys. Chem. B 2012, 116 (48) 14089– 14099 DOI: 10.1021/jp307782y108Influence of a Flavan-3-ol Substituent on the Affinity of Anthocyanins (Pigments) toward Vinylcatechin Dimers and Proanthocyanidins (Copigments)Nave, Frederico; Bras, Natercia F.; Cruz, Luis; Teixeira, Natercia; Mateus, Nuno; Ramos, Maria J.; Di Meo, Florent; Trouillas, Patrick; Dangles, Olivier; De Freitas, VictorJournal of Physical Chemistry B (2012), 116 (48), 14089-14099CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The aim of this study is to investigate interactions possibly taking place in red wine between three flavanols (copigments, CP), i.e., two epimeric vinylcatechin dimers (CP1 and CP2) and catechin dimer B3 (CP3), and a specific pigment resulting from the condensation between the main grape anthocyanin malvidin 3-O-glucoside (oenin) and catechin, catechin-(4→8)-oenin. By comparison with our previous work on oenin itself, the influence of the catechin moiety of the anthocyanin in the binding is established. The thermodn. parameters show that both vinylcatechin dimers exhibit a higher affinity for catechin-(4→8)-oenin, in comparison with proanthocyanidin B3, as previously obsd. with oenin. However, the corresponding binding consts. are weaker, probably due to steric hindrance in the anthocyanin brought by the flavanol nucleus. Consequently, catechin-(4→8)-oenin should be much less stabilized by copigmentation in hydroalcoholic soln. than oenin. Quantum mechanics and mol. dynamics simulations are also performed to interpret the binding data, to specify the relative arrangement of the pigment and copigment mols. within the complexes, and to interpret their absorption properties in the visible range.
- 109Dangles, O.; Elhajji, H. Synthesis of 3-Methoxy-and 3-B-D-Glucopyranosy1oxy) Flavylium Ions. Influence of the Flavylium Substitution Pattern on the Reactivity of Anthocyanins in Aqueous Solution Helv. Chim. Acta 1994, 77 (6) 1595– 1610 DOI: 10.1002/hlca.19940770616There is no corresponding record for this reference.
- 110Cruz, L.; Brás, N. F.; Teixeira, N.; Mateus, N.; Ramos, M. J.; Dangles, O.; De Freitas, V. Vinylcatechin Dimers Are Much Better Copigments for Anthocyanins than Catechin Dimer Procyanidin B3 J. Agric. Food Chem. 2010, 58 (5) 3159– 3166 DOI: 10.1021/jf9037419110Vinylcatechin Dimers Are Much Better Copigments for Anthocyanins than Catechin Dimer Procyanidin B3Cruz, Luis; Bras, Natercia F.; Teixeira, Natercia; Mateus, Nuno; Joao Ramos, Maria; Dangles, Olivier; De Freitas, VictorJournal of Agricultural and Food Chemistry (2010), 58 (5), 3159-3166CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The binding consts. (K) for the interaction of three copigments (CP), two epimeric vinylcatechin dimers (CP1 and CP2), and catechin dimer B3 (CP3) with two pigments, malvidin-3-glucoside (oenin) and malvidin-3,5-diglucoside (malvin), were detd. The K values clearly show that both vinylcatechin dimers have much higher affinity for oenin and malvin than dimer B3: KCP2 > KCP1 » KCP3. Quantum mechanics and mol. dynamics calcns. were also performed to interpret the binding data and specify the relative arrangement of the pigment and copigment mols. within the complexes.
- 111Malien-Aubert, C.; Dangles, O.; Amiot, M. J. Influence of Procyanidins on the Color Stability of Oenin Solutions J. Agric. Food Chem. 2002, 50 (11) 3299– 3305 DOI: 10.1021/jf011392b111Influence of Procyanidins on the Color Stability of Oenin SolutionsMalien-Aubert, Celine; Dangles, Olivier; Amiot, Marie JosepheJournal of Agricultural and Food Chemistry (2002), 50 (11), 3299-3305CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The aim of the present work was to specify the influence of the polymn. degree on the color stability of anthocyanins using model solns. under higher thermal conditions simulating rapid food aging. Results showed that an increase in polymeric degree improves the color stability of oenin. Solns. contg. a catechin tetramer, purified from brown rice, displayed a remarkable stability. Flavanols as monomers, (+)-catechin and (-)-epicatechin, appeared to decrease stability with the formation of a xanthylium salt leading to yellowish solns. For the dimers, procyanidin B2 and B3, different behaviors on red color stability have been obsd. corresponding to their different susceptibility to cleavage upon heating. In the presence of the trimeric procyanidin C2, the red color appeared more stable. However, the HPLC chromatograms showed a decrease in the amplitude of the peaks of oenin and procyanidin C2. Concomitantly, a new peak appeared with a maximal absorption in the red region. This newly formed pigment probably came from the condensation of oenin and procyanidin C2.
- 112Benesi, H. A.; Hildebrand, J. H. A Spectrophotometric Investigation of the Interaction of Iodine with Aromatic Hydrocarbons J. Am. Chem. Soc. 1949, 71 (8) 2703– 2707 DOI: 10.1021/ja01176a030112A spectrophotometric investigation of the interaction of iodine with aromatic hydrocarbonsBenesi, H. A.; Hildebrand, J. H.Journal of the American Chemical Society (1949), 71 (), 2703-7CODEN: JACSAT; ISSN:0002-7863.cf. C.A. 42, 8172g. The absorption spectra of I2 in C6H5CF3, benzene, toluene, o- and p-xylene, and mesitylene were measured in the region 270-700 mμ. In the visible region, the absorption peaks of these solns. showed moderate shifts toward shorter wave lengths in the order listed above. With the exception of I2 in C6H5CF3, each of the aromatic hydrocarbon solns. had an intense absorption band in the ultraviolet region which was shown to be characteristic of a complex contg. one I2 and one aromatic hydrocarbon mol. The equil. between I2 and the aromatic hydrocarbons was investigated in the neutral solvents CCl4 and C7H14, and the results show that the iodine-mesitylene complex is more stable than the iodine-benzene complex. These findings are strong evidence for an acid-base interaction between I2 and aromatic hydrocarbons. Absorption measurements also were made of I2 in CCl4, CS2, C7H14, Et2O, Me2CO, and 1,1-dichloroethane. No absorption bands analogous to those of the aromatic hydrocarbon solns. were found in the region 270-400 mμ.
- 113Dangles, O.; Brouillard, R. Polyphenol Interactions. The Copigmentation Case: Thermodynamic Data from Temperature Variation and Relaxation Kinetics. Medium Effect Can. J. Chem. 1992, 70 (8) 2174– 2189 DOI: 10.1139/v92-273There is no corresponding record for this reference.
- 114Di Meo, F.; Sancho Garcia, J. C.; Dangles, O.; Trouillas, P. Highlights on Anthocyanin Pigmentation and Copigmentation: A Matter of Flavonoid Π-Stacking Complexation To Be Described by DFT-D J. Chem. Theory Comput. 2012, 8 (6) 2034– 2043 DOI: 10.1021/ct300276p114Highlights on Anthocyanin Pigmentation and Copigmentation: A Matter of Flavonoid π-Stacking Complexation To Be Described by DFT-DDi Meo, Florent; Sancho Garcia, Juan Carlos; Dangles, Olivier; Trouillas, PatrickJournal of Chemical Theory and Computation (2012), 8 (6), 2034-2043CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Anthocyanidins are a class of π-conjugated systems responsible for red, blue, and purple colors of plants. They exhibit the capacity of aggregation in the presence of other natural compds. including flavonols. Such complexations induce color modulation in plants, which is known as copigmentation. It is largely driven by π-interactions existing between pigments and copigments. In this work, the energies of copigmentation-complexation and self-assocn. are systematically evaluated for an anthocyanidin/flavonol couple prototype (3-O-methylcyanidin/quercetin). To describe noncovalent interactions, DFT-D appears mandatory to reach a large accuracy. Due to the chem. complexity of this phenomenon, we also aim at assessing the relevance of both B3P86-D2 and ωB97X-D functionals. The benchmarking has shown that B3P86-D2 possesses enough accuracy when dealing with π-π interactions with respect to both spin component scaled Moller-Plesset second-order perturbation theory post Hartree-Fock method and exptl. data. UV-vis absorption properties are then evaluated with time-dependent DFT for the different complexes. The use of range-sepd. hybrid functionals, such as ωB97X-D, helped to correctly disentangle and interpret the origin of the UV-vis exptl. shifts attributed to the subtle copigmentation phenomenon.
- 115Awika, J. M. Behavior of 3-Deoxyanthocyanidins in the Presence of Phenolic Copigments Food Res. Int. 2008, 41 (5) 532– 538 DOI: 10.1016/j.foodres.2008.03.002115Behavior of 3-deoxyanthocyanidins in the presence of phenolic copigmentsAwika, Joseph M.Food Research International (2008), 41 (5), 532-538CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)Anthocyanin stability and color intensity are generally improved in the presence of copigments in moderately acidic environments. The 3-deoxyanthocyanins on the other hand are fairly stable to color loss due to change in pH. It is unknown, therefore, how they behave in the presence of copigments. We studied the effects of common phenolic copigments, tannic, ferulic, and O-coumaric acids, and rutin on behavior and stability of six 3-deoxyanthocyandins over 4.5 mo. Tannic and ferulic acid produced the most significant bathochromic shift, whereas rutin had no bathochromic effect. None of the copigments produced a significant hyperchromic shift with the pigments, implying colored species of the pigments were predominant under conditions used. Ferulic and tannic acids were the most effective at improving color stability of 5-hydroxylated pigments, whereas tannic acid and rutin improved the stability of 5-methoxylated pigments the most. Substitution at C-5 was key to overall behavior of the 3-deoxyanthocyanidins.
- 116El Hajji, H.; Dangles, O.; Figueiredo, P.; Brouillard, R. 3′-(β-D-Glycopyranosyloxy) Flavylium Ions: Synthesis and Investigation of Their Properties in Aqueous Solution. Hydrogen Bonding as a Mean of Colour Variation Helv. Chim. Acta 1997, 80 (2) 398– 413 DOI: 10.1002/hlca.19970800206There is no corresponding record for this reference.
- 117Limón, P. M.; Gavara, R.; Pina, F. Thermodynamics and Kinetics of Cyanidin 3-Glucoside and Caffeine Copigments J. Agric. Food Chem. 2013, 61 (22) 5245– 5251 DOI: 10.1021/jf4006643There is no corresponding record for this reference.
- 118Melo, M. J.; Moncada, M. C.; Pina, F. On the Red Colour of Raspberry (Rubus Idaeus) Tetrahedron Lett. 2000, 41 (12) 1987– 1991 DOI: 10.1016/S0040-4039(00)00080-0There is no corresponding record for this reference.
- 119Gierschner, J.; Lüer, L.; Milián-Medina, B.; Oelkrug, D.; Egelhaaf, H.-J. Highly Emissive H-Aggregates or Aggregation-Induced Emission Quenching? The Photophysics of All-Trans Para-Distyrylbenzene J. Phys. Chem. Lett. 2013, 4 (16) 2686– 2697 DOI: 10.1021/jz400985t119Highly Emissive H-Aggregates or Aggregation-Induced Emission Quenching? The Photophysics of All-Trans para-DistyrylbenzeneGierschner, Johannes; Luer, Larry; Milian-Medina, Begona; Oelkrug, Dieter; Egelhaaf, Hans-JoachimJournal of Physical Chemistry Letters (2013), 4 (16), 2686-2697CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The present Perspective critically reexamines the photophysics of para-distyrylbenzene (DSB) as a prototype of herringbone-arranged H-aggregates to resolve the apparent contradiction of the frequently reported aggregation-induced emission quenching in H-aggregates on one side and highly emissive DSB crystals on the other and discusses the signatures and fate of excitons in single- and polycryst. samples, including size and polarization effects.
- 120Gierschner, J.; Park, S. Y. Luminescent Distyrylbenzenes: Tailoring Molecular Structure and Crystalline Morphology J. Mater. Chem. C 2013, 1 (37) 5818– 5832 DOI: 10.1039/c3tc31062k120Luminescent distyrylbenzenes: tailoring molecular structure and crystalline morphologyGierschner, Johannes; Park, Soo YoungJournal of Materials Chemistry C: Materials for Optical and Electronic Devices (2013), 1 (37), 5818-5832CODEN: JMCCCX; ISSN:2050-7534. (Royal Society of Chemistry)A review. The last few years have seen a steady increase in small mol. based conjugated materials, which promise innovative (opto)electronic applications. This requires however a systematic understanding of structure-property relationships, which can only be achieved via libraries of structurally well-defined single cryst. materials based on systematically designed mol. structures. In this feature article, we are presenting structure-property relationships of functionalized distyrylbenzene (DSB), which is one of the most extensively investigated π-conjugated mol. materials. This will provide a general insight into the specific implications of intermol. arrangements on their solid state optoelectronic properties, discussing H- vs. J-aggregation, herringbone vs. π-stacks, the occurrence of excimers, size effects, and polycrystallinity. The systematic insight into DSB functionalization will then suggest pathways towards targeted mol. design strategies, with special focus on the cyano-vinylene motif (DCS materials) which allows for highly fluorescence solid state samples due to synergetic packing effects promoted by its twist elasticity and secondary bonding interaction. Finally, recent advances in the application of DSB-/DCS-based materials are shortly reviewed.
- 121Gonnet, J.-F. Colour Effects of Co-Pigmentation of Anthocyanins revisited—1. A Colorimetric Definition Using the CIELAB Scale Food Chem. 1998, 63 (3) 409– 415 DOI: 10.1016/S0308-8146(98)00053-3There is no corresponding record for this reference.
- 122Gordillo, B.; Rodríguez-Pulido, F. J.; Escudero-Gilete, M. L.; González-Miret, M. L.; Heredia, F. J. Comprehensive Colorimetric Study of Anthocyanic Copigmentation in Model Solutions. Effects of pH and Molar Ratio J. Agric. Food Chem. 2012, 60 (11) 2896– 2905 DOI: 10.1021/jf2046202122Comprehensive Colorimetric Study of Anthocyanic Copigmentation in Model Solutions. Effects of pH and Molar RatioGordillo, Belen; Rodriguez-Pulido, Francisco J.; Escudero-Gilete, M. Luisa; Gonzalez-Miret, M. Lourdes; Heredia, Francisco J.Journal of Agricultural and Food Chemistry (2012), 60 (11), 2896-2905CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)New colorimetric variables were defined in the uniform CIELAB color space to assess the quant. and qual. color changes induced by copigmentation and their incidence on visual perception. The copigmentation process was assayed in model solns. between malvidin 3-glucoside and 3 phenolic compds. (catechin, epicatechin, and caffeic acid) as a function of the pH and the pigment/copigment molar ratio. Along the pH variation, the greatest magnitude of copigmentation was obtained at pH 3.0, being significantly higher with epicatechin and caffeic acid. At high acidic pH, the main contribution of copigmentation to the total color was qual., whereas between pH 2.0 and 4.0, the main colorimetric contribution was quant. The contribution of epicatechin and caffeic acid to the color changes was more marked for the quant. characteristics. On contrast, particularly at higher pH values, the qual. contribution was more important in catechin copigmented solns. Increasing copigment concn. induced perceptible color changes at molar ratios higher than 1:2, consisting in a bluish and darkening effect of the anthocyanin solns. Among the different CIELAB attributes, hue difference was the best correlated parameter with the increase of copigment concn., proving the relevance of this physicochem. phenomenon on the qual. changes of anthocyanin color.
- 123Gonnet, J.-F. Colour Effects of Co-Pigmentation of Anthocyanins revisited—2.A Colorimetric Look at the Solutions of Cyanin Co-Pigmented Byrutin Using the CIELAB Scale Food Chem. 1999, 66 (3) 387– 394 DOI: 10.1016/S0308-8146(99)00088-6There is no corresponding record for this reference.
- 124Xu, H.; Liu, X.; Yan, Q.; Yuan, F.; Gao, Y. A Novel Copigment of Quercetagetin for Stabilization of Grape Skin Anthocyanins Food Chem. 2015, 166, 50– 55 DOI: 10.1016/j.foodchem.2014.05.125124A novel copigment of quercetagetin for stabilization of grape skin anthocyaninsXu, Honggao; Liu, Xuan; Yan, Qiuli; Yuan, Fang; Gao, YanxiangFood Chemistry (2015), 166 (), 50-55CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)The thermal and light stability of grape skin anthocyanins combined with quercetagetin was investigated at designed pH values of 3, 4 and 5. The molar ratios of anthocyanins to quercetagetin were 1:10, 1:20 and 1:40 for thermally treatment at 70 °C, 80 °C and 90 °C, resp., and the ratios were tested at 5:1, 1:1, 1:5 and 1:10 in the light exposure expts. The degrdn. reaction of anthocyanins in the presence of quercetagetin followed the first-order kinetic model. The half-life (t1/2) of anthocyanins was extended significantly with the increase of quercetagetin concn. (p < 0.05). The total color difference values (ΔE*) for the anthocyanin solns. with quercetagetin were smaller than those without copigment under the same exptl. conditions (pH and light exposure time). Compared with epigallocatechin gallate (EGCG), tea polyphenols (TP), myricitrin and rutin, quercetagetin was the most effective copigment to stabilize grape skin anthocyanins.
- 125Shao, P.; Zhang, J.; Fang, Z.; Sun, P. Complexing of Chlorogenic Acid with B-Cyclodextrins: Inclusion Effects, Antioxidative Properties and Potential Application in Grape Juice Food Hydrocolloids 2014, 41, 132– 139 DOI: 10.1016/j.foodhyd.2014.04.003There is no corresponding record for this reference.
- 126Mistry, T. V.; Cai, Y.; Lilley, T. H.; Haslam, E. Polyphenol Interactions. Part 5. Anthocyanin Co-Pigmentation J. Chem. Soc., Perkin Trans. 2 1991, 8) 1287– 1296 DOI: 10.1039/p29910001287126Polyphenol interactions. Part 5. Anthocyanin copigmentationMistry, Tarankumar V.; Cai, Ya; Lilley, Terence H.; Haslam, EdwinJournal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1991), (8), 1287-96CODEN: JCPKBH; ISSN:0300-9580.In aq. media, anthocyanins undergo several structural transformations and exist in a series of equil. between carbinol-base, flavylium cation, quinonoidal anhydro-base and chalcone forms. A detailed interpretation of the 1H-NMR (400 MHz) spectra of malvin in D2O is given for the first time in the context of these equil. The phenomenon of copigmentation is reviewed and the efficacy of various phenolic flavonoids, galloyl and hexahydroxycinnamyl esters as natural copigments is detd. quant. DNA, RNA and ATP also act as effective copigments for malvin in the form the flavylium ion but both caffeine and theophylline preferentially stabilize the quinonoidal-base forms of the anthocyanin. At pH 3.5 to 7.0, they give rise to stable violet through to blue and finally green colors. 1H-NMR studies of all these intermol. copigmentation reactions are reported and the phenomenon is provisionally interpreted in terms of hydrophobically reinforced π-π stacking of anthocyanin and copigment mols. in the aq. environment.
- 127Son, T.-D.; Chachaty, C. Nucleoside Conformations: XIV. Conformation of Adenosine Monophosphates in Aqueous Solution by Proton Magnetic Resonance Spectroscopy Biochim. Biophys. Acta, Nucleic Acids Protein Synth. 1974, 335 (1) 1– 13 DOI: 10.1016/0005-2787(74)90234-2There is no corresponding record for this reference.
- 128Leydet, Y.; Gavara, R.; Petrov, V.; Diniz, A. M.; Jorge Parola, A.; Lima, J. C.; Pina, F. The Effect of Self-Aggregation on the Determination of the Kinetic and Thermodynamic Constants of the Network of Chemical Reactions in 3-Glucoside Anthocyanins Phytochemistry 2012, 83, 125– 135 DOI: 10.1016/j.phytochem.2012.06.022128The effect of self-aggregation on the determination of the kinetic and thermodynamic constants of the network of chemical reactions in 3-glucoside anthocyaninsLeydet, Yoann; Gavara, Raquel; Petrov, Vesselin; Diniz, Ana M.; Parola, A. Jorge; Lima, Joao C.; Pina, FernandoPhytochemistry (Elsevier) (2012), 83 (), 125-135CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Ltd.)The six most common 3-glucoside anthocyanins, pelargonidin-3-glucoside, peonidin-3-glucoside, delphinidin-3-glucoside, malvidin-3-glucoside, cyanidin-3-glucoside and petunidin-3-glucoside were studied in great detail by NMR, UV-vis absorption and stopped flow. For each anthocyanin, the thermodn. and kinetic consts. of the network of chem. reactions were calcd. at different anthocyanin concn., from 6 × 10-6 M up to 8 × 10-4 M; an increasing of the flavylium cation acidity const. to give quinoidal base and a decreasing of the flavylium cation hydration const. to give hemiketal were obsd. by increasing the anthocyanin concn. These effects are attributed to the self-aggregation of the flavylium cation and quinoidal base, which is stronger in the last case. The UV-vis and 1H NMR spectral variations resulting from the increasing of the anthocyanin concn. were discussed in terms of two aggregation models; monomer-dimer and isodesmic, the last one considering the formation of higher order aggregates possessing the same aggregation const. of the dimer. The self-aggregation const. of flavylium cation at pH = 1.0, calcd. by both models increases by increasing the no. of methoxy (-OCH3) or hydroxy (-OH) substituents following the order: myrtillin (2 -OH), oenin (2 -OCH3), 3-OGl-petunidin (1 -OH, 1 -OCH3), kuromanin (1 -OH), 3-OGl-peonidin (1 -OCH3) and callistephin (none). Evidence for flavylium aggregates possessing a shape between J and H was achieved, as well as for the formation of higher order aggregates.
- 129Dimicoli, J. L.; Hélène, C. Complex Formation between Purine and Indole Derivatives in Aqueous Solutions. Proton Magnetic Resonance Studies J. Am. Chem. Soc. 1973, 95 (4) 1036– 1044 DOI: 10.1021/ja00785a008There is no corresponding record for this reference.
- 130Alluis, B.; Pérol, N.; El Hajji, H.; Dangles, O. Water-Soluble Flavonol (=3Hydroxy2-Phenyl-4H-1-Benzopyran-4-One) Derivatives: Chemical Synthesis, Colouring, and Antioxidant Properties Helv. Chim. Acta 2000, 83 (2) 428– 443 DOI: 10.1002/(SICI)1522-2675(20000216)83:2<428::AID-HLCA428>3.3.CO;2-AThere is no corresponding record for this reference.
- 131Hondo, T.; Yoshida, K.; Nakagawa, A.; Kawai, T.; Tamura, H.; Goto, T. Structural Basis of Blue-Colour Development in Flower Petals from Commelina Communis Nature 1992, 358 (6386) 515– 518 DOI: 10.1038/358515a0There is no corresponding record for this reference.
- 132Shiono, M.; Matsugaki, N.; Takeda, K. Phytochemistry: Structure of the Blue Cornflower Pigment Nature 2005, 436 (7052) 791– 791 DOI: 10.1038/436791aThere is no corresponding record for this reference.
- 133Dufour, C.; Sauvaitre, I. Interactions between Anthocyanins and Aroma Substances in a Model System. Effect on the Flavor of Grape-Derived Beverages J. Agric. Food Chem. 2000, 48 (5) 1784– 1788 DOI: 10.1021/jf990877l133Interactions between anthocyanins and aroma substances in a model system. Effect on the flavor of grape-derived beveragesDufour, Claire; Sauvaitre, IsabelleJournal of Agricultural and Food Chemistry (2000), 48 (5), 1784-1788CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Evaluation of the sensory quality of wine or grape-derived beverages led the authors to study the interactions between flavors and anthocyanins, the colored family of polyphenols. The flavylium cation-ligand complexation, resulting in copigmentation (rise in pigment visible absorption with a concomitant bathochromic shift), was investigated using visible absorption spectroscopy. Sole volatile phenols were found to markedly interact with malvidin-3,5-O-diglucoside. With series of guaiacyl-derived aroma substances, acyl-substituted ligands proved to be better copigments than alkyl-substituted ones. Assocn. consts. and 1:1 complex stoichiometry were further detd. for several substrates. Decreasing binding to malvin was obsd. for acetosyringone, syringaldehyde, acetovanillone, vanillin, 3,5-dimethoxyphenol, and 4-ethylguaiacol. Addn. of 10% ethanol lowered by one-third the assocn. consts. for malvin-ligand couples and for malvidin-3-O-glucoside with acetosyringone and syringaldehyde. The main driving force was ascribed to hydrophobicity, although this study evidenced an influence of the ligand substitution pattern on copigmentation.
- 134Tuominen, A.; Sinkkonen, J.; Karonen, M.; Salminen, J.-P. Sylvatiins, Acetylglucosylated Hydrolysable Tannins from the Petals of Geranium Sylvaticum Show Co-Pigment Effect Phytochemistry 2015, 115, 239– 251 DOI: 10.1016/j.phytochem.2015.01.005134Sylvatiins, acetylglucosylated hydrolysable tannins from the petals of Geranium sylvaticum show co-pigment effectTuominen, Anu; Sinkkonen, Jari; Karonen, Maarit; Salminen, Juha-PekkaPhytochemistry (Elsevier) (2015), 115 (), 239-251CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Ltd.)Four hydrolysable tannins, named as sylvatiins A (1), B (2), C (3) and D (4), were isolated from the petals of Geranium sylvaticum. On the basis of spectrometric evidence of NMR anal. (1H NMR, 13C NMR, DQF-COSY, TOCSY, NOESY, HSQC and HMBC), CD and ESI-MS/MS, sylvatiins A, B and C were characterized as galloyl glucoses contg. one or two acetylglucoses attached to the 3-OH of the galloyl group, whereas sylvatiin D was found to have a chebulinic acid core contg. acetylglucose attached in a similar way. The potential of these compds. to act as defensive compds. against herbivores was evaluated using the radial diffusion assay that measures the protein pptn. capacity. In addn., the capacity of sylvatiins to act as co-pigments with anthocyanins of G. sylvaticum petals was measured in vitro at different pH values. Sylvatiins A and D maintained efficiently the purple flower color near the natural pH of petal cells. The amt. of sylvatiins was changed according to the flower color; deep purple petals with higher amt. of anthocyanin contained more sylvatiins A and C than whiter petals. It was concluded that G. sylvaticum petal cells may accumulate sylvatiins for intermol. co-pigmentation purposes.
- 135Dangles, O.; Saito, N.; Brouillard, R. Kinetic and Thermodynamic Control of Flavylium Hydration in the Pelargonidin-Cinnamic Acid Complexation. Origin of the Extraordinary Flower Color Diversity of Pharbitis Nil J. Am. Chem. Soc. 1993, 115 (8) 3125– 3132 DOI: 10.1021/ja00061a011There is no corresponding record for this reference.
- 136Eiro, M. J.; Heinonen, M. Anthocyanin Color Behavior and Stability during Storage: Effect of Intermolecular Copigmentation J. Agric. Food Chem. 2002, 50 (25) 7461– 7466 DOI: 10.1021/jf0258306136Anthocyanin Color Behavior and Stability during Storage: Effect of Intermolecular CopigmentationEiro, Maarit J.; Heinonen, MarinaJournal of Agricultural and Food Chemistry (2002), 50 (25), 7461-7466CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Intermol. copigmentation reactions are significantly responsible for the manifold color expression of fruits, berries, and their products. These reactions were investigated with 5 anthocyanins and 5 phenolic acids acting as copigments. The stability of the pigment-copigment complexes formed was studied during a storage period of 6 mo. The study was conducted using a UV-visible spectrophotometer to monitor the hyperchromic effect and the bathochromic shift of the complexes. The greatest copigmentation reactions took place in malvidin 3-glucoside solns. The strongest copigments for all anthocyanins were ferulic and rosmarinic acids. The immediate reaction of rosmarinic acid with malvidin 3-glucoside resulted in the biggest bathochromic shift (19 nm) and the strongest hyperchromic effect, increasing the color intensity by 260%. The color induced by rosmarinic acid was not very stable. The color intensity of pelargonidin 3-glucoside increased greatly throughout the storage period with the addn. of ferulic and caffeic acids.
- 137Houbiers, C.; Lima, J. C.; Maçanita, A. L.; Santos, H. Color Stabilization of Malvidin 3-Glucoside: Self-Aggregation of the Flavylium Cation and Copigmentation with the Z-Chalcone Form J. Phys. Chem. B 1998, 102 (18) 3578– 3585 DOI: 10.1021/jp972320jThere is no corresponding record for this reference.
- 138Elhabiri, M.; Figueiredo, P.; Toki, K.; Figueiredo, P.; Toki, K.; Saito, N.; Brouillard, R.; Figueiredo, P.; Toki, K. Anthocyanin–aluminium and – gallium Complexes in Aqueoussolution J. Chem. Soc., Perkin Trans. 2 1997, 2) 355– 362 DOI: 10.1039/a603851d138Anthocyanin-aluminum and -gallium complexes in aqueous solutionElhabiri, M.; Figueiredo, P.; Toki, K.; Saito, N.; Brouillard, R.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1997), (2), 355-362CODEN: JCPKBH; ISSN:0300-9580. (Royal Society of Chemistry)Complexation of aluminum and gallium ions with synthetic anthocyanin models and natural anthocyanins extd. from the blue flowers of Evolvulus pilosus cv 'Blue Daze' and the violet flowers of Matthiola incana has been thoroughly investigated in aq. soln. From UV-VIS spectroscopic data collected at pH 2-5, the presence of complexes, involving not only the colored forms but also the colorless forms of the pigments is demonstrated. A theor. treatment is developed for the calcn. of the corresponding stability consts. The pigments studied throughout this work can be divided into two series, one sharing a cyanidin chromophore and the other a delphinidin one. Within both series, individual pigments are distinguished according to the degree and type of glycosidation and/or acylation. Intramol. effects such as copigmentation of anthocyanin-aluminum complexes and the effect of the presence of a malonyl group on the formation of those complexes are discussed. These results are important to plant pigmentation and, for instance, a narrow pH domain in which color amplification due to complexation is at a max. has been found.
- 139Al Bittar, S.; Mora, N.; Loonis, M.; Dangles, O. Chemically Synthesized Glycosides of Hydroxylated Flavylium Ions as Suitable Models of Anthocyanins: Binding to Iron Ions and Human Serum Albumin, Antioxidant Activity in Model Gastric Conditions Molecules 2014, 19 (12) 20709– 20730 DOI: 10.3390/molecules191220709There is no corresponding record for this reference.
- 140Kasha, M.; Rawls, H. R.; Ashraf El-Bayoumi, M. The Exciton Model in Molecular Spectroscopy Pure Appl. Chem. 1965, 11 (3–4) 371– 392 DOI: 10.1351/pac196511030371140Exciton model in molecular spectroscopyKasha, Michael; Rawls, Henry R.; El-Bayoumi, M. AshrafPure and Applied Chemistry (1965), 11 (3-4), 371-92CODEN: PACHAS; ISSN:0033-4545.A summary is presented of the application of the mol. exciton model to dimers, trimers, and double and triple mols. In those cases where no significant exciton effect is observable in the singlet-singlet absorption spectrum for the composite mol., the enhancement of lowest triplet state excitation may still be conspicuous and significant.
- 141Yoshida, K.; Toyama, Y.; Kameda, K.; Kondo, T. Contribution of Each Caffeoyl Residue of the Pigment Molecule of Gentiodelphin to Blue Color Development Phytochemistry 2000, 54 (1) 85– 92 DOI: 10.1016/S0031-9422(00)00049-2There is no corresponding record for this reference.
- 142Yoshida, K.; Kondo, T.; Goto, T. Intramolecular Stacking Conformation of Gentiodelphin, a Diacylated Anthocyanin from Gentiana Makinoi Tetrahedron 1992, 48 (21) 4313– 4326 DOI: 10.1016/S0040-4020(01)80442-7There is no corresponding record for this reference.
- 143Nerdal, W.; Andersen, Ø. M. Evidence for Self-Association of the Anthocyanin Petanin in Acidified, Methanolic Solution Using Two-Dimensional Nuclear Overhauser Enhancement NMR Experiments and Distance Geometry Calculations Phytochem. Anal. 1991, 2 (6) 263– 270 DOI: 10.1002/pca.2800020606143Evidence for self-association of the anthocyanin petanin in acidified, methanolic solution using two-dimensional nuclear Overhauser enhancement NMR experiments and distance geometry calculationsNerdal, Willy; Andersen, Oeyvind M.Phytochemical Analysis (1991), 2 (6), 263-70CODEN: PHANEL; ISSN:0958-0344.A new mechanism is described for self-assocn. of anthocyanins in soln. Petanin [petunidin 3-O-[6-O-(4-O-E-p-coumaroyl-α-L-rhamnopyranosyl)-β-D-glucopyranoside]-5-O-β-D-glucopuranoside] was studied in acidified methanolic soln. using the nuclear Overhauser enhancement (NOESY) NMR technique. Intra- and inter-mol. NOESY cross-peaks were obsd., and the corresponding proton-proton distance bounds were used in distance geometry calcns. to det. the stacking of the petanin aglycons. The orientation of two self-assocd. petanin aglycons was found to be head-to-tail along both the long and the short aglycon axis. Lack of obsd. NOESY cross-peaks between protons of the coumaroyl group and the aglycon indicated absence of intramol. stacking of the stable petanin mols. Non-coplanarity between the planes of the benzopyrylium and the Ph rings was also indicated.
- 144Nerdal, W.; Andersen, Ø. M. Intermolecular Aromatic Acid Association of an Anthocyanin (petanin) Evidenced by Two-Dimensional Nuclear Overhauser Enhancement Nuclear Magnetic Resonance Experiments and Distance Geometry Calculations Phytochem. Anal. 1992, 3 (4) 182– 189 DOI: 10.1002/pca.2800030408144Intermolecular aromatic acid association of an anthocyanin (petanin) evidence by two-dimensional nuclear Overhauser enhancement nuclear magnetic resonance experiments and distance geometry calculationsNerdal, Willy; Andersen, Oeyvind M.Phytochemical Analysis (1992), 3 (4), 182-9CODEN: PHANEL; ISSN:0958-0344.A new mechanism for the mol. assocn. of the arom acid of an anthocyanin using two-dimensional 1H NOESY expts., is described.
- 145Yoshida, K.; Kondo, T.; Goto, T. Unusually Stable Monoacylated Anthocyanin from Purple Yam Dioscorea Alata Tetrahedron Lett. 1991, 32 (40) 5579– 5580 DOI: 10.1016/0040-4039(91)80088-NThere is no corresponding record for this reference.
- 146Mori, M.; Miki, N.; Ito, D.; Kondo, T.; Yoshida, K. Structure of Tecophilin, a Tri-Caffeoylanthocyanin from the Blue Petals of Tecophilaea Cyanocrocus, and the Mechanism of Blue Color Development Tetrahedron 2014, 70 (45) 8657– 8664 DOI: 10.1016/j.tet.2014.09.046146Structure of tecophilin, a tri-caffeoylanthocyanin from the blue petals of Tecophilaea cyanocrocus, and the mechanism of blue color developmentMori, Mihoko; Miki, Naoko; Ito, Daisuke; Kondo, Tadao; Yoshida, KumiTetrahedron (2014), 70 (45), 8657-8664CODEN: TETRAB; ISSN:0040-4020. (Elsevier Ltd.)The pigment, tecophilin, in blue flowers of Tecophilaea cyanocrocus was isolated and the structure was detd. to be 3-O-(6-O-α-L-rhamnopyranosyl-β-D-glucopyranosyl)-7-O-(6-O-(4-O-(2-O-(4-O-β-D-glucopyranosyl-(E)-caffeoyl)-6-O-(4-O-β-D-glucopyranosyl-(E)-caffeoyl)-β-D-glucopyranosyl)-(E)-caffeoyl)-β-D-glucopyranosyl)delphinidin. The reprodn. expt. of the same color as petals according to the results of chem. anal. and measurement of vacuolar pH of blue cells clarified that the blue color solely develops by tecophilin without interaction of metal ions nor co-pigments. 1H NMR anal. and CD spectrum indicate the co-existence of clockwise intermol. self-assocn. of the delphinidin nuclei and intramol. π-π stacking between the chromophore and caffeoyl residues to derive bathochromic shift of the absorption spectrum and stabilize the color by preventing hydration reaction.
- 147Terahara, N.; Callebaut, A.; Ohba, R.; Nagata, T.; Ohnishi-Kameyama, M.; Suzuki, M. Triacylated Anthocyanins from Ajuga Reptans Flowers and Cell Cultures Phytochemistry 1996, 42 (1) 199– 203 DOI: 10.1016/0031-9422(95)00838-1There is no corresponding record for this reference.
- 148Terahara, N.; Toki, K.; Saito, N.; Honda, T.; Matsui, T.; Osajima, Y. Eight New Anthocyanins, Ternatins C1-C5 and D3 and Preternatins A3 and C4 from Young Clitoria Ternatea Flowers J. Nat. Prod. 1998, 61 (11) 1361– 1367 DOI: 10.1021/np980160cThere is no corresponding record for this reference.
- 149Escribano-Bailón, T.; Dangles, O.; Brouillard, R. Coupling Reactions between Flavylium Ions and Catechin Phytochemistry 1996, 41 (6) 1583– 1592 DOI: 10.1016/0031-9422(95)00811-X149Coupling reactions between flavylium ions and catechinEscribano-Bailon, Teresa; Dangles, Olivier; Brouillard, RaymondPhytochemistry (1996), 41 (6), 1583-92CODEN: PYTCAS; ISSN:0031-9422. (Elsevier)In order to model natural polymeric pigments present in old red wines, new covalent adducts have been synthesized upon condensation of synthetic flavylium ions (models of anthocyanins) with catechin (model of tannins) in the presence and in the absence of acetaldehyde. These new pigments have been investigated by 1D and 2D NMR, HPLC, FAB-mass and UV-visible spectroscopies and mol. modeling. The two flavylium salts used in this work (3,4'-dimethoxy-7-hydroxyflavylium chloride and 5,7-dihydroxy-3,4'-dimethoxyflavylium chloride) display quite different reactivities toward catechin. The electronic donating effect of the catechin moiety and the formation of noncovalent dimers in acidic aq. or methanolic soln. should be mainly responsible for the improved stability of the flavylium chromophore in the new pigments.
- 150Dueñas, M.; Fulcrand, H.; Cheynier, V. Formation of Anthocyanin–flavanol Adducts in Model Solutions Anal. Chim. Acta 2006, 563 (1–2) 15– 25 DOI: 10.1016/j.aca.2005.10.062There is no corresponding record for this reference.
- 151Dueñas, M.; Salas, E.; Cheynier, V.; Dangles, O.; Fulcrand, H. UV–Visible Spectroscopic Investigation of the 8,8-Methylmethine Catechin-Malvidin 3-Glucoside Pigments in Aqueous Solution: Structural Transformations and Molecular Complexation with Chlorogenic Acid J. Agric. Food Chem. 2006, 54 (1) 189– 196 DOI: 10.1021/jf0516989There is no corresponding record for this reference.
- 152Chassaing, S.; Lefeuvre, D.; Jacquet, R.; Jourdes, M.; Ducasse, L.; Galland, S.; Grelard, A.; Saucier, C.; Teissedre, P.-L.; Dangles, O. Physicochemical Studies of New Anthocyano-Ellagitannin Hybrid Pigments: About the Origin of the Influence of Oak C-Glycosidic Ellagitannins on Wine Color Eur. J. Org. Chem. 2010, 2010 (1) 55– 63 DOI: 10.1002/ejoc.200901133There is no corresponding record for this reference.
- 153Bloor, S. J. Overview of Methods for Analysis and Identification of Flavonoids Methods Enzymol. 2001, 335, 3– 14 DOI: 10.1016/S0076-6879(01)35227-8There is no corresponding record for this reference.
- 154Dangles, O.; Saito, N.; Brouillard, R. Anthocyanin Intramolecular Copigment Effect Phytochemistry 1993, 34 (1) 119– 124 DOI: 10.1016/S0031-9422(00)90792-1There is no corresponding record for this reference.
- 155Figueiredo, P.; George, F.; Tatsuzawa, F.; Toki, K.; Saito, N.; Brouillard, R. New Features of Intramolecular Copigmentation by Acylated Anthocyanins Phytochemistry 1999, 51, 125– 132 DOI: 10.1016/S0031-9422(98)00685-2155New features of intramolecular copigmentation by acylated anthocyaninsFigueiredo, Paulo; George, Florian; Tatsuzawa, Fumi; Toki, Kenjiro; Saito, Norio; Brouillard, RaymondPhytochemistry (1999), 51 (1), 125-132CODEN: PYTCAS; ISSN:0031-9422. (Elsevier Science Ltd.)Three series of structurally related anthocyanins, extd. from the red-purple flowers of Dendrobium "Pramot", xLaeliocattleya cv. Mini Purple, Bletilla striata and Phalaenopsis all belonging to the Orchidaceae family and another series extd. from the pink flowers of Senecio cruentus (Compositae) allowed the confirmation of the existence of strong intramol. copigmentation effects. These interactions confer stability to the colored forms of the mols., in a wide range of slightly acidic to neutral aq. media. Moreover, the existence of structural relationships among the four series stressed the different influences exerted by the diverse substituent groups. The existence of a malonylglucoside attached to position 3 of all but three of the mols. put forward a new role for the malonyl residue, in this particular position.
- 156Fleschhut, J.; Kratzer, F.; Rechkemmer, G.; Kulling, S. E. Stability and Biotransformation of Various Dietary Anthocyanins in Vitro Eur. J. Nutr. 2006, 45 (1) 7– 18 DOI: 10.1007/s00394-005-0557-8156Stability and biotransformation of various dietary anthocyanins in vitroFleschhut, Jens; Kratzer, Frank; Rechkemmer, Gerhard; Kulling, Sabine E.European Journal of Nutrition (2006), 45 (1), 7-18CODEN: EJNUFZ; ISSN:1436-6207. (Steinkopff Verlag)Background Anthocyanins, which are found in high concns. in fruit and vegetable, may play a beneficial role in retarding or reversing the course of chronic degenerative diseases. However, little is known about the biotransformation and the metab. of anthocyanins so far. Aim of the study The aim of the study was to investigate possible transformation pathways of anthocyanins by human fecal microflora and by rat liver microsomes as a source of cytochrome P 450 enzymes as well as of glucuronyltransferases. Methods Pure anthocyanins, an aq. ext. of red radish as well as the assumed degrdn. products were incubated with human fecal suspension. The incubation mixts. were purified by solid-phase extn. and analyzed by HPLC/DAD/MS and GC/MS. Quantification was done by the external std. method. Furthermore the biotransformation of anthocyanins by incubation with rat liver microsomes in the presence of the cofactor NADPH (as a model for the phase I oxidn.) and in the presence of activated glucuronic acid (as a model for the phase II glucuronidation) was investigated. Results Glycosylated and acylated anthocyanins were rapidly degraded by the intestinal microflora after anaerobic incubation with a human fecal suspension. The major stable products of anthocyanin degrdn. are the corresponding phenolic acids derived from the B-ring of the anthocyanin skeleton. Anthocyanins were not metabolized by cytochrome P 450 enzymes, neither hydroxylated nor demethylated. However they were glucuronidated by rat liver microsomes to several products. Conclusions The gut microflora seem to play an important role in the biotransformation of anthocyanins. A rapid degrdn. could be one major reason for the poor bioavailability of anthocyanins in pharmacokinetic studies described so far in the literature. The formation of phenolic acids as the major stable degrdn. products gives an important hint to the fate of anthocyanins in vivo.
- 157Lopes, P.; Richard, T.; Saucier, C.; Teissedre, P.-L.; Monti, J.-P.; Glories, Y. Anthocyanone A: A Quinone Methide Derivative Resulting from Malvidin 3- O -Glucoside Degradation J. Agric. Food Chem. 2007, 55 (7) 2698– 2704 DOI: 10.1021/jf062875oThere is no corresponding record for this reference.
- 158Sadilova, E.; Carle, R.; Stintzing, F. C. Thermal Degradation of Anthocyanins and Its Impact on Color Andin Vitro Antioxidant Capacity Mol. Nutr. Food Res. 2007, 51 (12) 1461– 1471 DOI: 10.1002/mnfr.200700179158Thermal degradation of anthocyanins and its impact on color and in vitro antioxidant capacitySadilova, Eva; Carle, Reinhold; Stintzing, Florian C.Molecular Nutrition & Food Research (2007), 51 (12), 1461-1471CODEN: MNFRCV; ISSN:1613-4125. (Wiley-VCH Verlag GmbH & Co. KGaA)The aim of the current study was to thoroughly investigate the structural changes of anthocyanins at pH 3.5 in purified fractions from black carrot, elderberry and strawberry heated over 6 h at 95°C. Degrdn. products were monitored by HPLC-DAD-MS3 to elucidate the prevailing degrdn. pathways. In addn., alterations of color and antioxidant properties obsd. upon heating were scrutinized. Most interestingly, the degrdn. pathways at pH 3.5 were found to differ from those at pH 1. Among others, chalcone glycosides were detected at 320 nm in heat-treated elderberry and strawberry pigment isolates, and opening of the pyrylium ring initiated anthocyanin degrdn. In the case of acylated anthocyanins, acyl-glycoside moieties were split off from the flavylium backbone, first. Finally, for all pigment isolates, phenolic acids and phloroglucinaldehyde were the terminal degrdn. products as remainders of the B- and A-ring, resp. Maximum and min. antioxidant stabilizing capacities were found in black carrot and strawberry, resp., which was explained by the high degree of acylation in the former. After heating, decline of trolox equiv. antioxidant capacity (TEAC) was obsd. in all samples, which was attributed to both anthocyanins and their colorless degrdn. products following thermal exposure. As deduced from the ratio of TEAC value and anthocyanin content, the loss of anthocyanin bioactivity could not be compensated by the antioxidant capacity of newly formed colorless phenolics upon heating.
- 159Yang, J.; Hu, W.; Usvyat, D.; Matthews, D.; Schütz, M.; Chan, G. K.-L. Ab Initio Determination of the Crystalline Benzene Lattice Energy to Sub-Kilojoule/mol Accuracy Science 2014, 345 (6197) 640– 643 DOI: 10.1126/science.1254419159Ab initio determination of the crystalline benzene lattice energy to sub-kilojoule/mole accuracyYang, Jun; Hu, Weifeng; Usvyat, Denis; Matthews, Devin; Schuetz, Martin; Chan, Garnet Kin-LicScience (Washington, DC, United States) (2014), 345 (6197), 640-643CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Computation of lattice energies to an accuracy sufficient to distinguish polymorphs is a fundamental bottleneck in crystal structure prediction. For the lattice energy of the prototypical benzene crystal, we combined the quantum chem. advances of the last decade to attain sub-kilojoule per mol accuracy, an order-of-magnitude improvement in certainty over prior calcns. that necessitates revision of the exptl. extrapolation to 0 K. Our computations reveal the nature of binding by improving on previously inaccessible or inaccurate multibody and many-electron contributions and provide revised ests. of the effects of temp., vibrations, and relaxation. Our demonstration raises prospects for definitive first-principles resoln. of competing polymorphs in mol. crystal structure prediction.
- 160Cavalcanti, R. N.; Santos, D. T.; Meireles, M. A. A. Non-Thermal Stabilization Mechanisms of Anthocyanins in Model and Food systems—An Overview Food Res. Int. 2011, 44 (2) 499– 509 DOI: 10.1016/j.foodres.2010.12.007160Non-thermal stabilization mechanisms of anthocyanins in model and food systems-An overviewCavalcanti, Rodrigo N.; Santos, Diego T.; Meireles, Maria Angela A.Food Research International (2011), 44 (2), 499-509CODEN: FORIEU; ISSN:0963-9969. (Elsevier B.V.)A review. Phenolic compds. are part of the secondary metab. of plants and are of great importance for their survival in unfavorable environment. A class of phenolic compds. easily found in the Plant Kingdom, is anthocyanins, a flavonoid category. They are water-sol. pigments that confer the bright red, blue, and purple colors of fruits and vegetables and promote several health benefits due to their diverse biol. activities. Different factors affect the color and stability of these compds. including pH, temp., light, presence of copigments, self-assocn., metallic ions, enzymes, oxygen, ascorbic acid, sugar, among others. For this reason many studies have been conducted with the aim to increase the stability of these substances. Therefore, the present review highlights studies on the stabilization of anthocyanins and presents latent anthocyanin stabilization mechanisms and demonstrates the potentiality of the main techniques used: assocn. and encapsulation.
- 161Rodrigues, R. F.; Ferreira da Silva, P.; Shimizu, K.; Freitas, A. A.; Kovalenko, S. A.; Ernsting, N. P.; Quina, F. H.; Maçanita, A. Ultrafast Internal Conversion in a Model Anthocyanin-Polyphenol Complex: Implications for the Biological Role of Anthocyanins in Vegetative Tissues of Plants Chem. - Eur. J. 2009, 15 (6) 1397– 1402 DOI: 10.1002/chem.200801207161Ultrafast internal conversion in a model anthocyanin-polyphenol complex: implications for the biological role of anthocyanins in vegetative tissues of plantsRodrigues, Rita Franca; Ferreira da Silva, Palmira; Shimizu, Karina; Freitas, Adilson A.; Kovalenko, Sergey A.; Ernsting, Nikolaus P.; Quina, Frank H.; Macanita, AntonioChemistry - A European Journal (2009), 15 (6), 1397-1402CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The red flavylium cations of anthocyanins form ground-state charge-transfer complexes with several naturally occurring electron-donor copigments, such as hydroxylated flavones and hydroxycinnamic or benzoic acids. Excitation of the 7-methoxy-4-methyl-flavylium-protocatechuic acid complex results in ultrafast (240 fs) internal conversion to the ground state of the complex by way of a low-lying charge-transfer state. Thus, both uncomplexed anthocyanins, whose excited state decays by fast (5-20 ps) excited-state proton transfer, and anthocyanin-copigment complexes have highly efficient mechanisms of deactivation that are consistent with the proposed protective role of anthocyanins against excess solar radiation in the vegetative tissues of plants.
- 162Ferreira da Silva, P.; Paulo, L.; Barbafina, A.; Elisei, F.; Quina, F. H.; Maçanita, A. L. Photoprotection and the Photophysics of Acylated Anthocyanins Chem. - Eur. J. 2012, 18 (12) 3736– 3744 DOI: 10.1002/chem.201102247162Photoprotection and the Photophysics of Acylated AnthocyaninsFerreira da Silva, Palmira; Paulo, Luisa; Barbafina, Adrianna; Eisei, Fausto; Quina, Frank H.; Macanita, Antonio L.Chemistry - A European Journal (2012), 18 (12), 3736-3744CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The proposed role of anthocyanins in protecting plants against excess solar radiation is consistent with the occurrence of ultrafast (5-25 ps) excited-state proton transfer as the major deexcitation pathway of these mols. However, because natural anthocyanins absorb mainly in the visible region of the spectra, with only a narrow absorption band in the UV-B region, this highly efficient deactivation mechanism would essentially only protect the plant from visible light. On the other hand, ground-state charge-transfer complexes of anthocyanins with naturally occurring electron-donor copigments, such as hydroxylated flavones, flavonoids, and hydroxycinnamic or benzoic acids, do exhibit high UV-B absorptivities that complement that of the anthocyanins. In this work, we report a comparative study of the photophysics of the naturally occurring anthocyanin cyanin, intermol. cyanin-coumaric acid complexes, and an acylated anthocyanin, i.e., cyanin with a pendant coumaric ester copigment. Both inter- and intramol. anthocyanin-copigment complexes are shown to have ultrafast energy dissipation pathways comparable to those of model flavylium cation-copigment complexes. However, from the standpoint of photoprotection, the results indicate that the covalent attachment of copigment mols. to the anthocyanin represents a much more efficient strategy by providing the plant with significant UV-B absorption capacity and at the same time coupling this absorption to efficient energy dissipation pathways (ultrafast internal conversion of the complexed form and fast energy transfer from the excited copigment to the anthocyanin followed by adiabatic proton transfer) that avoid net photochem. damage.
- 163Song, B. J.; Sapper, T. N.; Burtch, C. E.; Brimmer, K.; Goldschmidt, M.; Ferruzzi, M. G. Photo- and Thermodegradation of Anthocyanins from Grape and Purple Sweet Potato in Model Beverage Systems J. Agric. Food Chem. 2013, 61 (6) 1364– 1372 DOI: 10.1021/jf3044007There is no corresponding record for this reference.
- 164Malien-Aubert, C.; Dangles, O.; Amiot, M. J. Color Stability of Commercial Anthocyanin-Based Extracts in Relation to the Phenolic Composition. Protective Effects by Intra- and Intermolecular Copigmentation J. Agric. Food Chem. 2001, 49 (1) 170– 176 DOI: 10.1021/jf000791oThere is no corresponding record for this reference.
- 165Cabrita, L.; Petrov, V.; Pina, F. On the Thermal Degradation of Anthocyanidins: Cyanidin RSC Adv. 2014, 4 (36) 18939 DOI: 10.1039/c3ra47809b165On the thermal degradation of anthocyanidins: cyanidinCabrita, Luis; Petrov, Vesselin; Pina, FernandoRSC Advances (2014), 4 (36), 18939-18944CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Cyanidin was studied by direct pH jumps (from equilibrated solns. at very low pH values to higher pH values) and reverse pH jumps (from equilibrated or not equilibrated solns. at higher pH values to very low ones). The kinetic steps of the direct and reverse pH jumps were followed by stopped flow, absorption spectroscopy and HPLC, at different timescales. The pH dependent rate const. of the slower kinetic process to reach the equil. follows a bell shaped curve as described for many synthetic flavylium compds. Unlike anthocyanins, it was proved that there is no pH dependent reversibility in the system, since the chalcone suffers an irreversible degrdn. process. The math. expression to describe the bell shaped behavior was deduced. These results contribute to explain why in plants glycosylation is crucial for the stabilization of the anthocyanins.
- 166Schneider, H.-J. Dispersive Interactions in Solution Complexes Acc. Chem. Res. 2015, 48 (7) 1815– 1822 DOI: 10.1021/acs.accounts.5b00111166Dispersive Interactions in Solution ComplexesSchneider, Hans-JoergAccounts of Chemical Research (2015), 48 (7), 1815-1822CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Dispersive interactions are known to play a major role in mol. assocns. in the gas phase and in the solid state. In soln., however, their significance has been disputed in recent years on the basis of several arguments. A major problem until now has been the sepn. of dispersive and hydrophobic effects, which are both maximized in water due the low polarizability of this most important medium. Analyses of complexes between porphyrins and systematically varied substrates in water have allowed us to discriminate dispersive from hydrophobic effects, as the latter turned out to be negligible for complexations with flat surfaces such as porphyrins. Also, for the first time, it has become possible to obtain binding free energy increments ΔΔG for a multitude of org. residues including halogen, amide, amino, ether, carbonyl, ester, nitro, sulfur, unsatured, and cyclopropane groups, which turned out to be additive. Binding contributions for satd. residues are unmeasurably small, with ΔΔG > 1 kJ/mol, but they increase to, e.g., ΔΔG = 5 kJ/mol for a nitro group, a value not far from, e.g., that of a stacking pyridine ring. Stacking interactions of heteroarenes with porphyrins depend essentially on the size of the arenes, in line with polarizabilities, and seem to be rather independent of the position of nitrogen within the rings. Measurements of halogen derivs. indicate that complexes with porphyrins, cyclodextrins, and pillarenes as hosts in different media consistently show increasing stability from fluorine to iodine as the substituent. This, and the obsd. sequence with other substrates, is in line with the expected increase in dispersive forces with increasing polarizability. Induced dipoles, which also would increase with polarizability, can be ruled out as providing the driving source in view of the data with halides: the obsd. stability sequence is opposite the change of electronegativity from fluorine to iodine. The same holds for the solvent effect obsd. in ethanol-water mixts. Dispersive contributions vary not only with the polarizability of the used media but also with the interacting receptor sites; it has been shown that for cucurbiturils the polarizability inside the cavity is extremely low, which also explains why hydrophobic effects are maximized with these hosts. Complexations with other known host compds., however, such as those between cryptands or cavitands with, e.g., noble gases, bear the signature of dominating dispersive forces. Some recent examples illustrate that such van der Waals forces can also play an important role in complexations with proteins. Again, a clue for this is the increase in ΔG for inhibitor binding by 7 kJ/mol for, e.g., a bromine in comparison to a fluorine deriv.
- 167Hunter, C. A.; Sanders, J. K. M. The Nature of.pi.-.pi. Interactions J. Am. Chem. Soc. 1990, 112 (14) 5525– 5534 DOI: 10.1021/ja00170a016167The nature of π-π interactionsHunter, Christopher A.; Sanders, Jeremy K. M.Journal of the American Chemical Society (1990), 112 (14), 5525-34CODEN: JACSAT; ISSN:0002-7863.A simple model of the charge distribution in a π-system is used to explain the strong geometrical requirements for interactions between arom. mols. The key feature of the model is that it considers the σ-framework and the π-electrons sep. and demonstrates that net favorable π-π interactions are actually the result of π-σ attractions that overcome π-π repulsions. The calcns. correlate with observations made on porphyrin π-π interactions both in soln. and in the cryst. state. By using an idealized π-atom, some general rules for predicting the geometry of favorable π-π interactions are derived. In particular a favorable offset or slipped geometry is predicted. These rules successfully predict the geometry of intermol. interactions in the crystal structures of arom. mols. and rationalize a range of host-guest phenomena. The theory demonstrates that the electron donor-acceptor concept can be misleading: it is the properties of the atoms at the points of intermol. contact rather than the overall mol. properties which are important.
- 168Hwang, J.; Li, P.; Carroll, W. R.; Smith, M. D.; Pellechia, P. J.; Shimizu, K. D. Additivity of Substituent Effects in Aromatic Stacking Interactions J. Am. Chem. Soc. 2014, 136 (40) 14060– 14067 DOI: 10.1021/ja504378p168Additivity of Substituent Effects in Aromatic Stacking InteractionsHwang, Jungwun; Li, Ping; Carroll, William R.; Smith, Mark D.; Pellechia, Perry J.; Shimizu, Ken D.Journal of the American Chemical Society (2014), 136 (40), 14060-14067CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The goal of this study was to exptl. test the additivity of the electrostatic substituent effects (SEs) for the arom. stacking interaction. The additivity of the SEs was assessed using a small mol. model system that could adopt an offset face-to-face arom. stacking geometry. The intramol. interactions of these mol. torsional balances were quant. measured via the changes in a folded/unfolded conformational equil. Five different types of substituents were examd. (CH3, OCH3, Cl, CN, and NO2) that ranged from electron-donating to electron-withdrawing. The strength of the intramol. stacking interactions was measured for 21 substituted arom. stacking balances and 21 control balances in chloroform soln. The obsd. stability trends were consistent with additive SEs. Specifically, additive SE models could predict SEs with an accuracy from ±0.01 to ±0.02 kcal/mol. The additive SEs were consistent with Wheeler and Houk's direct SE model. However, the indirect or polarization SE model cannot be ruled out as it shows similar levels of additivity for two to three substituent systems, which were the no. of substituents in our model system. SE additivity also has practical utility as the SEs can be accurately predicted. This should aid in the rational design and optimization of systems that utilize arom. stacking interactions.
- 169Cockroft, S. L.; Hunter, C. A.; Lawson, K. R.; Perkins, J.; Urch, C. J. Electrostatic Control of Aromatic Stacking Interactions J. Am. Chem. Soc. 2005, 127 (24) 8594– 8595 DOI: 10.1021/ja050880n169Electrostatic Control of Aromatic Stacking InteractionsCockroft, Scott L.; Hunter, Christopher A.; Lawson, Kevin R.; Perkins, Julie; Urch, Christopher J.Journal of the American Chemical Society (2005), 127 (24), 8594-8595CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A supramol. approach has been used to investigate the free energies of intermol. arom. stacking interactions. Chem. double mutant cycles have been used to measure the effect of a range of substituents on face-to-face stacking interactions with Ph and pentafluorophenyl rings. Electrostatic effects dominate the trends in interaction energy.
- 170Meyer, E. A.; Castellano, R. K.; Diederich, F. Interactions with Aromatic Rings in Chemical and Biological Recognition Angew. Chem., Int. Ed. 2003, 42 (11) 1210– 1250 DOI: 10.1002/anie.200390319170Interactions with aromatic rings in chemical and biological recognitionMeyer, Emmanuel A.; Castellano, Ronald K.; Diederich, FrancoisAngewandte Chemie, International Edition (2003), 42 (11), 1210-1250CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Intermol. interactions involving arom. rings are key processes in both chem. and biol. recognition. Their understanding is essential for rational drug design and lead optimization in medicinal chem. Different approaches-biol. studies, mol. recognition studies with artificial receptors, crystallog. database mining, gas-phase studies, and theor. calcns. - are pursued to generate a profound understanding of the structural and energetic parameters of individual recognition modes involving arom. rings. This review attempts to combine and summarize the knowledge gained from these investigations. The review focuses mainly on examples with biol. relevance since one of its aims it to enhance the knowledge of mol. recognition forces that is essential for drug development.
- 171Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic Rings in Chemical and Biological Recognition: Energetics and Structures Angew. Chem., Int. Ed. 2011, 50 (21) 4808– 4842 DOI: 10.1002/anie.201007560171Aromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.
- 172Hunter, C. A. Quantifying Intermolecular Interactions: Guidelines for the Molecular Recognition Toolbox Angew. Chem., Int. Ed. 2004, 43 (40) 5310– 5324 DOI: 10.1002/anie.200301739172Quantifying intermolecular interactions: Guidelines for the molecular recognition toolboxHunter, Christopher A.Angewandte Chemie, International Edition (2004), 43 (40), 5310-5324CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Mol. recognition events in soln. are affected by many different factors that have hampered the development of an understanding of intermol. interactions at a quant. level. Tendency is to partition these effects into discrete phenomenol. fields that are classified, named, and divorced: arom. interactions, cation-T interactions, CH-O hydrogen bonds, short strong hydrogen bonds, and hydrophobic interactions to name a few. To progress in the field, the authors need to develop an integrated quant. appreciation of the relative magnitudes of all of the different effects that might influence the mol. recognition behavior of a given system. In an effort to navigate undergraduates through the vast and sometimes contradictory literature on the subject, I have developed an approach that treats theor. ideas and exptl. observations about intermol. interactions in the gas phase, the solid state, and soln. from a single simplistic viewpoint. The key features are outlined here, and although many of the ideas will be familiar, the aim is to provide a semiquant. thermodn. ranking of these effects in soln. at room temp.
- 173Trouillas, P.; Di Meo, F.; Gierschner, J.; Linares, M.; Sancho-García, J. C.; Otyepka, M. Optical Properties of Wine Pigments: Theoretical Guidelines with New Methodological Perspectives Tetrahedron 2015, 71 (20) 3079– 3088 DOI: 10.1016/j.tet.2014.10.046There is no corresponding record for this reference.
- 174London, F. The General Theory of Molecular Forces Trans. Faraday Soc. 1937, 33 (0) 8b– 26 DOI: 10.1039/tf937330008bThere is no corresponding record for this reference.
- 175Chandler, D. Interfaces and the Driving Force of Hydrophobic Assembly Nature 2005, 437 (7059) 640– 647 DOI: 10.1038/nature04162175Interfaces and the driving force of hydrophobic assemblyChandler, DavidNature (London, United Kingdom) (2005), 437 (7059), 640-647CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. The hydrophobic effect - the tendency for oil and water to segregate - is important in diverse phenomena, from the cleaning of laundry, to the creation of micro-emulsions to make new materials, to the assembly of proteins into functional complexes. This effect is multifaceted depending on whether hydrophobic mols. are individually hydrated or driven to assemble into larger structures. Despite the basic principles underlying the hydrophobic effect being qual. well understood, only recently have theor. developments begun to explain and quantify many features of this ubiquitous phenomenon.
- 176Biedermann, F.; Nau, W. M.; Schneider, H.-J. The Hydrophobic Effect Revisited—Studies with Supramolecular Complexes Imply High-Energy Water as a Noncovalent Driving Force Angew. Chem., Int. Ed. 2014, 53 (42) 11158– 11171 DOI: 10.1002/anie.201310958176The Hydrophobic Effect Revisited-Studies with Supramolecular Complexes Imply High-Energy Water as a Noncovalent Driving ForceBiedermann, Frank; Nau, Werner M.; Schneider, Hans-JoergAngewandte Chemie, International Edition (2014), 53 (42), 11158-11171CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Traditional descriptions of the hydrophobic effect from entropic arguments or the calcn. of solvent-occupied surfaces must be questioned in view of new results obtained with supramol. complexes. In these studies, it was possible to sep. hydrophobic from dispersive interactions, which are strongest in aq. systems. Even very hydrophobic alkanes assoc. significantly only in cavities contg. water mols. with an insufficient no. of possible hydrogen bonds. The replacement of high-energy water in cavities by guest mols. is the essential enthalpic driving force for complexation, as borne out by data for complexes of cyclodextrins, cyclophanes, and cucurbiturils, for which complexation enthalpies of up to -100 kJ mol-1 were reached for encapsulated alkyl residues. Water-box simulations were used to characterize the different contributions from high-energy water and enabled the calcn. of the assocn. free enthalpies for selected cucurbituril complexes to within a 10% deviation from exptl. values. Cavities in artificial receptors are more apt to show the enthalpic effect of high-energy water than those in proteins or nucleic acids, because they bear fewer or no functional groups in the inner cavity to stabilize interior water mols.
- 177Waters, M. L. Aromatic Interactions Acc. Chem. Res. 2013, 46 (4) 873– 873 DOI: 10.1021/ar4000828177Aromatic InteractionsWaters, Marcey L.Accounts of Chemical Research (2013), 46 (4), 873CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)There is no expanded citation for this reference.
- 178Yang, L.; Adam, C.; Nichol, G. S.; Cockroft, S. L. How Much Do van Der Waals Dispersion Forces Contribute to Molecular Recognition in Solution? Nat. Chem. 2013, 5 (12) 1006– 1010 DOI: 10.1038/nchem.1779178How much do van der Waals dispersion forces contribute to molecular recognition in solution?Yang, Lixu; Adam, Catherine; Nichol, Gary S.; Cockroft, Scott L.Nature Chemistry (2013), 5 (12), 1006-1010CODEN: NCAHBB; ISSN:1755-4330. (Nature Publishing Group)The emergent properties that arise from self-assembly and mol. recognition phenomena are a direct consequence of noncovalent interactions. Gas-phase measurements and computational methods point to the dominance of dispersion forces in mol. assocn., but solvent effects complicate the unambiguous quantification of these forces in soln. Here, the authors used synthetic mol. balances to measure interactions between apolar alkyl chains in 31 org., fluorous and aq. solvent environments. The exptl. interaction energies are an order of magnitude smaller than ests. of dispersion forces between alkyl chains that were derived from vaporization enthalpies and dispersion-cor. calcns. Instead, cohesive solvent-solvent interactions are the major driving force behind apolar assocn. in soln. Probably theor. models that implicate important roles for dispersion forces in mol. recognition events should be interpreted with caution in solvent-accessible systems.
- 179Mackerell, A. D. Empirical Force Fields for Biological Macromolecules: Overview and Issues J. Comput. Chem. 2004, 25 (13) 1584– 1604 DOI: 10.1002/jcc.20082179Empirical force fields for biological macromolecules: Overview and issuesMacKerell, Alexander D., Jr.Journal of Computational Chemistry (2004), 25 (13), 1584-1604CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A review. Empirical force field-based studies of biol. macromols. are becoming a common tool for investigating their structure-activity relationships at an at. level of detail. Such studies facilitate interpretation of exptl. data and allow for information not readily accessible to exptl. methods to be obtained. A large part of the success of empirical force field-based methods is the quality of the force fields combined with the algorithmic advances that allow for more accurate reprodn. of exptl. observables. Presented is an overview of the issues assocd. with the development and application of empirical force fields to biomol. systems. This is followed by a summary of the force fields commonly applied to the different classes of biomols.; proteins, nucleic acids, lipids, and carbohydrates. In addn., issues assocd. with computational studies on "heterogeneous" biomol. systems and the transferability of force fields to a wide range of org. mols. of pharmacol. interest are discussed.
- 180Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field J. Comput. Chem. 2004, 25 (9) 1157– 1174 DOI: 10.1002/jcc.20035180Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 181Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic Molecules J. Am. Chem. Soc. 1996, 118 (9) 2309– 2309 DOI: 10.1021/ja955032e181A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic Molecules [ Erratum for J. Am. Chem. Soc. 1995, 117, 5179-5197 ]Cornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M. Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1996), 118 (9), 2309CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)There is no expanded citation for this reference.
- 182Zgarbová, M.; Otyepka, M.; Sponer, J.; Hobza, P.; Jurecka, P. Large-Scale Compensation of Errors in Pairwise-Additive Empirical Force Fields: Comparison of AMBER Intermolecular Terms with Rigorous DFT-SAPT Calculations Phys. Chem. Chem. Phys. 2010, 12 (35) 10476– 10493 DOI: 10.1039/c002656eThere is no corresponding record for this reference.
- 183Šponer, J.; Banáš, P.; Jurečka, P.; Zgarbová, M.; Kührová, P.; Havrila, M.; Krepl, M.; Stadlbauer, P.; Otyepka, M. Molecular Dynamics Simulations of Nucleic Acids. From Tetranucleotides to the Ribosome J. Phys. Chem. Lett. 2014, 5 (10) 1771– 1782 DOI: 10.1021/jz500557y183Molecular Dynamics Simulations of Nucleic Acids. From Tetranucleotides to the RibosomeSponer, Jiri; Banas, Pavel; Jurecka, Petr; Zgarbova, Marie; Kuhrova, Petra; Havrila, Marek; Krepl, Miroslav; Stadlbauer, Petr; Otyepka, MichalJournal of Physical Chemistry Letters (2014), 5 (10), 1771-1782CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)A review. We present a brief overview of explicit solvent mol. dynamics (MD) simulations of nucleic acids. We explain phys. chem. limitations of the simulations, namely, the mol. mechanics (MM) force field (FF) approxn. and limited time scale. Further, we discuss relations and differences between simulations and expts., compare std. and enhanced sampling simulations, discuss the role of starting structures, comment on different versions of nucleic acid FFs, and relate MM computations with contemporary quantum chem. Despite its limitations, we show that MD is a powerful technique for studying the structural dynamics of nucleic acids with a fast growing potential that substantially complements exptl. results and aids their interpretation.
- 184Adcock, S. A.; McCammon, J. A. Molecular Dynamics: Survey of Methods for Simulating the Activity of Proteins Chem. Rev. 2006, 106 (5) 1589– 1615 DOI: 10.1021/cr040426m184Molecular dynamics: Survey of methods for simulating the activity of proteinsAdcock, Stewart A.; McCammon, J. AndrewChemical Reviews (Washington, DC, United States) (2006), 106 (5), 1589-1615CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Mol. dynamics simulations (MDS) of proteins have provided many insights into the internal motions of these biomols. Simulation of in silico models aids in the interpretation and reconciliation of exptl. data. With ongoing advances in both methodol. and computational resources, MDS are being extended to larger systems and longer time scales. This enables the investigation of motions and conformational changes that have functional implications and yields information that is not available though any other means. Today's results suggest that (subject to the continuing utilization of synergies between expt. and simulation) the applications of MDS will command an increasingly crit. role in the understanding of biol. systems.
- 185Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. The Amber Biomolecular Simulation Programs J. Comput. Chem. 2005, 26 (16) 1668– 1688 DOI: 10.1002/jcc.20290185The amber biomolecular simulation programsCase, David A.; Cheatham, Thomas E., III; Darden, Tom; Gohlke, Holger; Luo, Ray; Merz, Kenneth M., Jr.; Onufriev, Alexey; Simmerling, Carlos; Wang, Bing; Woods, Robert J.Journal of Computational Chemistry (2005), 26 (16), 1668-1688CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The authors describe the development, current features, and some directions for future development of the Amber package of computer programs. This package evolved from a program that was constructed in the late 1970s to do Assisted Model Building with Energy Refinement, and now contains a group of programs embodying a no. of powerful tools of modern computational chem., focused on mol. dynamics and free energy calcns. of proteins, nucleic acids, and carbohydrates.
- 186Brooks, B. R.; Brooks, C. L.; MacKerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S. CHARMM: The Biomolecular Simulation Program J. Comput. Chem. 2009, 30 (10) 1545– 1614 DOI: 10.1002/jcc.21287186CHARMM: The biomolecular simulation programBrooks, B. R.; Brooks, C. L., III; Mackerell, A. D., Jr.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; Yang, W.; York, D. M.; Karplus, M.Journal of Computational Chemistry (2009), 30 (10), 1545-1614CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A review. CHARMM (Chem. at HARvard Mol. Mechanics) is a highly versatile and widely used mol. simulation program. It has been developed over the last three decades with a primary focus on mols. of biol. interest, including proteins, peptides, lipids, nucleic acids, carbohydrates, and small mol. ligands, as they occur in soln., crystals, and membrane environments. For the study of such systems, the program provides a large suite of computational tools that include numerous conformational and path sampling methods, free energy estimators, mol. minimization, dynamics, and anal. techniques, and model-building capabilities. The CHARMM program is applicable to problems involving a much broader class of many-particle systems. Calcns. with CHARMM can be performed using a no. of different energy functions and models, from mixed quantum mech.-mol. mech. force fields, to all-atom classical potential energy functions with explicit solvent and various boundary conditions, to implicit solvent and membrane models. The program has been ported to numerous platforms in both serial and parallel architectures. This article provides an overview of the program as it exists today with an emphasis on developments since the publication of the original CHARMM article in 1983. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009.
- 187Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD J. Comput. Chem. 2005, 26 (16) 1781– 1802 DOI: 10.1002/jcc.20289187Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 188Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C. GROMACS: Fast, Flexible, and Free J. Comput. Chem. 2005, 26 (16) 1701– 1718 DOI: 10.1002/jcc.20291188GROMACS: Fast, flexible, and freeVan Der Spoel, David; Lindahl, Erik; Hess, Berk; Groenhof, Gerrit; Mark, Alan E.; Berendsen, Herman J. C.Journal of Computational Chemistry (2005), 26 (16), 1701-1718CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)This article describes the software suite GROMACS (Groningen MAchine for Chem. Simulation) that was developed at the University of Groningen, The Netherlands, in the early 1990s. The software, written in ANSI C, originates from a parallel hardware project, and is well suited for parallelization on processor clusters. By careful optimization of neighbor searching and of inner loop performance, GROMACS is a very fast program for mol. dynamics simulation. It does not have a force field of its own, but is compatible with GROMOS, OPLS, AMBER, and ENCAD force fields. In addn., it can handle polarizable shell models and flexible constraints. The program is versatile, as force routines can be added by the user, tabulated functions can be specified, and analyses can be easily customized. Nonequil. dynamics and free energy detns. are incorporated. Interfaces with popular quantum-chem. packages (MOPAC, GAMES-UK, GAUSSIAN) are provided to perform mixed MM/QM simulations. The package includes about 100 utility and anal. programs. GROMACS is in the public domain and distributed (with source code and documentation) under the GNU General Public License. It is maintained by a group of developers from the Universities of Groningen, Uppsala, and Stockholm, and the Max Planck Institute for Polymer Research in Mainz. Its Web site is http://www.gromacs.org.
- 189Leach, A. R. Molecular Modelling: Principles and Applications; Pearson, 2001.There is no corresponding record for this reference.
- 190Vashisth, H.; Skiniotis, G.; Brooks, C. L. Collective Variable Approaches for Single Molecule Flexible Fitting and Enhanced Sampling Chem. Rev. 2014, 114 (6) 3353– 3365 DOI: 10.1021/cr4005988There is no corresponding record for this reference.
- 191Bruccoleri, R. E.; Karplus, M. Conformational Sampling Using High-Temperature Molecular Dynamics Biopolymers 1990, 29 (14) 1847– 1862 DOI: 10.1002/bip.360291415There is no corresponding record for this reference.
- 192Chocholoušová, J.; Vacek, J.; Hobza, P. Acetic Acid Dimer in the Gas Phase, Nonpolar Solvent, Microhydrated Environment, and Dilute and Concentrated Acetic Acid: Ab Initio Quantum Chemical and Molecular Dynamics Simulations J. Phys. Chem. A 2003, 107 (17) 3086– 3092 DOI: 10.1021/jp027637k192Acetic Acid Dimer in the Gas Phase, Nonpolar Solvent, Microhydrated Environment, and Dilute and Concentrated Acetic Acid: Ab Initio Quantum Chemical and Molecular Dynamics SimulationsChocholousova, Jana; Vacek, Jaroslav; Hobza, PavelJournal of Physical Chemistry A (2003), 107 (17), 3086-3092CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Theor. study of the acetic acid dimer, its microhydration and its behavior in water and chloroform soln. was performed. To characterize the system, the authors adopted ab initio methods at the DFT and RI MP2 (the resoln. of the identity approxn. MP2) levels for the gas-phase calcns., PCM (polarizable continuum model) approxn. using the polarizable conductor calcn. model (COSMO) for description of solvent, and const. energy (NVE) and const. temp. (NVT) mol. dynamics simulations for gas phase and explicit solvent calcns., resp. The cyclic structure of the acetic acid dimer is the most stable in the gas phase only. During microhydration, the water mols. are incorporated in the dimer leading to water-sepd. structures. This conclusion is based on ab initio quantum chem. calcns., as well as on mol. dynamics simulations. The fact that the cyclic structure does not appear in water soln. is in agreement with previous theor. and exptl. results. Extending the search also on other acetic acid dimer structures, acetic acid does not form any dimer structure in water soln. The cyclic structure also probably is stable in chloroform soln.
- 193Zelený, T.; Hobza, P.; Kabelác, M. Microhydration of Guanine···cytosine Base Pairs, a Theoretical Study on the Role of Water in Stability, Structure and Tautomeric Equilibrium Phys. Chem. Chem. Phys. 2009, 11 (18) 3430– 3435 DOI: 10.1039/b819350aThere is no corresponding record for this reference.
- 194Sugita, Y.; Okamoto, Y. Replica-Exchange Molecular Dynamics Method for Protein Folding Chem. Phys. Lett. 1999, 314 (1–2) 141– 151 DOI: 10.1016/S0009-2614(99)01123-9194Replica-exchange molecular dynamics method for protein foldingSugita, Y.; Okamoto, Y.Chemical Physics Letters (1999), 314 (1,2), 141-151CODEN: CHPLBC; ISSN:0009-2614. (Elsevier Science B.V.)We have developed a formulation for mol. dynamics algorithm for the replica-exchange method. The effectiveness of the method for the protein-folding problem is tested with the penta-peptide Met-enkephalin. The method can overcome the multiple-min. problem by exchanging non-interacting replicas of the system at several temps. From only one simulation run, one can obtain probability distributions in canonical ensemble for a wide temp. range using multiple-histogram re-weighting techniques, which allows the calcn. of any thermodn. quantity as a function of temp. in that range.
- 195Beck, D. A. C.; White, G. W. N.; Daggett, V. Exploring the Energy Landscape of Protein Folding Using Replica-Exchange and Conventional Molecular Dynamics Simulations J. Struct. Biol. 2007, 157 (3) 514– 523 DOI: 10.1016/j.jsb.2006.10.002There is no corresponding record for this reference.
- 196Laio, A.; Parrinello, M. Escaping Free-Energy Minima Proc. Natl. Acad. Sci. U. S. A. 2002, 99 (20) 12562– 12566 DOI: 10.1073/pnas.202427399196Escaping free-energy minimaLaio, Alessandro; Parrinello, MicheleProceedings of the National Academy of Sciences of the United States of America (2002), 99 (20), 12562-12566CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)We introduce a powerful method for exploring the properties of the multidimensional free energy surfaces (FESs) of complex many-body systems by means of coarse-grained non-Markovian dynamics in the space defined by a few collective coordinates. A characteristic feature of these dynamics is the presence of a history-dependent potential term that, in time, fills the min. in the FES, allowing the efficient exploration and accurate detn. of the FES as a function of the collective coordinates. We demonstrate the usefulness of this approach in the case of the dissocn. of a NaCl mol. in water and in the study of the conformational changes of a dialanine in soln.
- 197Barducci, A.; Bonomi, M.; Parrinello, M. Metadynamics Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1 (5) 826– 843 DOI: 10.1002/wcms.31197MetadynamicsBarducci, Alessandro; Bonomi, Massimiliano; Parrinello, MicheleWiley Interdisciplinary Reviews: Computational Molecular Science (2011), 1 (5), 826-843CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Metadynamics is a powerful technique for enhancing sampling in mol. dynamics simulations and reconstructing the free-energy surface as a function of few selected degrees of freedom, often referred to as collective variables (CVs). In metadynamics, sampling is accelerated by a history-dependent bias potential, which is adaptively constructed in the space of the CVs. Since its first appearance, significant improvements have been made to the original algorithm, leading to an efficient, flexible, and accurate method that has found many successful applications in several domains of science. Here, we discuss first the theory underlying metadynamics and its recent developments. In particular, we focus on the crucial issue of choosing an appropriate set of CVs and on the possible strategies to alleviate this difficulty. Later in the second part, we present a few recent representative applications, which we have classified into three main classes: solid-state physics, chem. reactions, and biomols.
- 198Brás, N. F.; Cruz, L.; Fernandes, P. A.; De Freitas, V.; Ramos, M. J. Conformational Study of Two Diasteroisomers of Vinylcatechin Dimers in a Methanol Solution Int. J. Quantum Chem. 2011, 111 (7–8) 1498– 1510 DOI: 10.1002/qua.22631There is no corresponding record for this reference.
- 199Kollman, P. A.; Massova, I.; Reyes, C.; Kuhn, B.; Huo, S.; Chong, L.; Lee, M.; Lee, T.; Duan, Y.; Wang, W. Calculating Structures and Free Energies of Complex Molecules: Combining Molecular Mechanics and Continuum Models Acc. Chem. Res. 2000, 33 (12) 889– 897 DOI: 10.1021/ar000033j199Calculating Structures and Free Energies of Complex Molecules: Combining Molecular Mechanics and Continuum ModelsKollman, Peter A.; Massova, Irina; Reyes, Carolina; Kuhn, Bernd; Huo, Shuanghong; Chong, Lillian; Lee, Matthew; Lee, Taisung; Duan, Yong; Wang, Wei; Donini, Oreola; Cieplak, Piotr; Srinivasan, Jaysharee; Case, David A.; Cheatham, Thomas E., IIIAccounts of Chemical Research (2000), 33 (12), 889-897CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review, with 63 refs. A historical perspective on the application of mol. dynamics (MD) to biol. macromols. is presented. Recent developments combining state-of-the-art force fields with continuum solvation calcns. have allowed us to reach the fourth era of MD applications in which one can often derive both accurate structure and accurate relative free energies from mol. dynamics trajectories. We illustrate such applications on nucleic acid duplexes, RNA hairpins, protein folding trajectories, and protein-ligand, protein-protein, and protein-nucleic acid interactions.
- 200Bahar, I.; Lezon, T. R.; Bakan, A.; Shrivastava, I. H. Normal Mode Analysis of Biomolecular Structures: Functional Mechanisms of Membrane Proteins Chem. Rev. 2010, 110 (3) 1463– 1497 DOI: 10.1021/cr900095e200Normal Mode Analysis of Biomolecular Structures: Functional Mechanisms of Membrane ProteinsBahar, Ivet; Lezon, Timothy R.; Bakan, Ahmet; Shrivastava, Indira H.Chemical Reviews (Washington, DC, United States) (2010), 110 (3), 1463-1497CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Structural dynamics and function of membrane proteins are discussed in relation to principal component anal. of exptl. resolved conformations, normal mode anal. and elastic network models.
- 201Andricioaei, I.; Karplus, M. On the Calculation of Entropy from Covariance Matrices of the Atomic Fluctuations J. Chem. Phys. 2001, 115 (14) 6289– 6292 DOI: 10.1063/1.1401821201On the calculation of entropy from covariance matrices of the atomic fluctuationsAndricioaei, Ioan; Karplus, MartinJournal of Chemical Physics (2001), 115 (14), 6289-6292CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)An ad hoc method for calcg. the entropy of a biomol. system from the covariance matrix of the at. fluctuations is analyzed. It is shown that its essential assumption can be eliminated by a quasiharmonic anal. The computer time required for use of the latter is of the same order as that of the former.
- 202Tsui, V.; Case, D. A. Theory and Applications of the Generalized Born Solvation Model in Macromolecular Simulations Biopolymers 2000, 56 (4) 275– 291 DOI: 10.1002/1097-0282(2000)56:4<275::AID-BIP10024>3.0.CO;2-E202Theory and applications of the generalized Born solvation model in macromolecular simulationsTsui V; Case D ABiopolymers (2000-2001), 56 (4), 275-91 ISSN:0006-3525.Generalized Born (GB) models provide an attractive way to include some thermodynamic aspects of aqueous solvation into simulations that do not explicitly model the solvent molecules. Here we discuss our recent experience with this model, presenting in detail the way it is implemented and parallelized in the AMBER molecular modeling code. We compare results using the GB model (or GB plus a surface-area based "hydrophobic" term) to explicit solvent simulations for a 10 base-pair DNA oligomer, and for the 108-residue protein thioredoxin. A slight modification of our earlier suggested parameters makes the GB results more like those found in explicit solvent, primarily by slightly increasing the strength of NH [bond] O and NH [bond] N internal hydrogen bonds. Timing and energy stability results are reported, with an eye toward using these model for simulations of larger macromolecular systems and longer time scales.
- 203Rocchia, W.; Alexov, E.; Honig, B. Extending the Applicability of the Nonlinear Poisson–Boltzmann Equation: Multiple Dielectric Constants and Multivalent Ions J. Phys. Chem. B 2001, 105 (28) 6507– 6514 DOI: 10.1021/jp010454y203Extending the applicability of the nonlinear Poisson-Boltzmann equation: multiple dielectric constants and multivalent ionsRocchia, W.; Alexov, E.; Honig, B.Journal of Physical Chemistry B (2001), 105 (28), 6507-6514CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A new version of the DelPhi program, which provides numerical solns. to the nonlinear Poisson-Boltzmann (PB) equation, is reported. The program can divide space into multiple regions contg. different dielec. consts. and can treat systems contg. mixed salt solns. where the valence and concn. of each ion is different. The electrostatic free energy is calcd. by decompg. the various energy terms into Coulombic interactions so that that the calcd. free energies are independent of the lattice used to solve the PB equation. This, together with algorithms that optimally position polarization charges on the mol. surface, leads to a significant decrease in the dependence of the electrostatic free energy on the resoln. of the lattice used to solve the PB equation and, hence, to a remarkable improvement in the precision of the calcd. values. The Gauss-Seidel algorithm used in the current version of DelPhi is retained so that the new program retains many of the optimization features of the old one. The program uses dynamic memory allocation and can easily handle systems requiring large grid dimensions - for example, a 3003 system can be conveniently treated on a single SGI R12000 processor. An algorithm that ests. the best relaxation parameter to solve the nonlinear equation for a given system is described, and is implemented in the program at run time. A no. of applications of the program are presented.
- 204Baker, C. M.; Lopes, P. E. M.; Zhu, X.; Roux, B.; MacKerell, A. D. Accurate Calculation of Hydration Free Energies Using Pair-Specific Lennard-Jones Parameters in the CHARMM Drude Polarizable Force Field J. Chem. Theory Comput. 2010, 6 (4) 1181– 1198 DOI: 10.1021/ct9005773204Accurate Calculation of Hydration Free Energies using Pair-Specific Lennard-Jones Parameters in the CHARMM Drude Polarizable Force FieldBaker, Christopher M.; Lopes, Pedro E. M.; Zhu, Xiao; Roux, Benoit; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2010), 6 (4), 1181-1198CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Lennard-Jones (LJ) parameters for a variety of model compds. have previously been optimized within the CHARMM Drude polarizable force field to reproduce accurately pure liq. phase thermodn. properties as well as addnl. target data. While the polarizable force field resulting from this optimization procedure has been shown to satisfactorily reproduce a wide range of exptl. ref. data across numerous series of small mols., a slight but systematic overestimate of the hydration free energies has also been noted. Here, the reprodn. of exptl. hydration free energies is greatly improved by the introduction of pair-specific LJ parameters between solute heavy atoms and water oxygen atoms that override the std. LJ parameters obtained from combining rules. The changes are small and a systematic protocol is developed for the optimization of pair-specific LJ parameters and applied to the development of pair-specific LJ parameters for alkanes, alcs., and ethers. The resulting parameters not only yield hydration free energies in good agreement with exptl. values, but also provide a framework upon which other pair-specific LJ parameters can be added as new compds. are parametrized within the CHARMM Drude polarizable force field. Detailed anal. of the contributions to the hydration free energies reveals that the dispersion interaction is the main source of the systematic errors in the hydration free energies. This information suggests that the systematic error may result from problems with the LJ combining rules and is combined with anal. of the pair-specific LJ parameters obtained in this work to identify a preliminary improved combining rule.
- 205Allinger, N. L. Conformational Analysis. 130. MM2. A Hydrocarbon Force Field Utilizing V1 and V2 Torsional Terms J. Am. Chem. Soc. 1977, 99 (25) 8127– 8134 DOI: 10.1021/ja00467a001205Conformational analysis. 130. MM2. A hydrocarbon force field utilizing V1 and V2 torsional termsAllinger, Norman L.Journal of the American Chemical Society (1977), 99 (25), 8127-34CODEN: JACSAT; ISSN:0002-7863.An improved force field for mol.-mechanics calcns. of the structures and energies of hydrocarbons is presented. The problem of simultaneously obtaining a sufficiently large gauche butane interaction energy while keeping the H atoms small enough for good structural predictions was solved with 1- and 2-fold rotational barriers. For 42 selected diverse types of hydrocarbons, the std. deviation between the calcd. and exptl. heats of formation is 0.42 kcal/mol, compared with an av. reported exptl. error for the same group of compds. of 0.40 kcal/mol.
- 206Charlton, A. J.; Davis, A. L.; Jones, D. P.; Lewis, J. R.; Davies, A. P.; Haslam, E.; Williamson, M. P. The Self-Association of the Black Tea Polyphenol Theaflavin and Its Complexation with Caffeine J. Chem. Soc. Perkin Trans. 2 2000, 2) 317– 322 DOI: 10.1039/a906380cThere is no corresponding record for this reference.
- 207Kunsági-Máté, S.; Szabó, K.; Nikfardjam, M. P.; Kollár, L. Determination of the Thermodynamic Parameters of the Complex Formation between Malvidin-3-O-Glucoside and Polyphenols. Copigmentation Effect in Red Wines J. Biochem. Biophys. Methods 2006, 69 (1–2) 113– 119 DOI: 10.1016/j.jbbm.2006.03.014207Determination of the thermodynamic parameters of the complex formation between malvidin-3-O-glucoside and polyphenols. Copigmentation effect in red winesKunsagi-Mate, Sandor; Szabo, Kornelia; Nikfardjam, Martin P.; Kollar, LaszloJournal of Biochemical and Biophysical Methods (2006), 69 (1-2), 113-119CODEN: JBBMDG; ISSN:0165-022X. (Elsevier Ltd.)The thermodn. of the mol. assocn. process between the malvidin-3-O-glucoside and a series of polyphenol derivs. (called copigmentation' in food chem.) were studied in aq. media. The Gibbs free energy, enthalpy and entropy values were detd. by the fluorometric method. A combination of the Job's method with the van't Hoff theory was applied for data evaluation. The results show the exothermic character of the copigmentation process. The change of the enthalpy seems to be the same in every complexation step. However, the decreasing of the entropy term is higher at higher stoichiometries. As a result, the Gibbs free energy changes and, thus, the complex stability decreases quickly with increasing stoichiometry. Quantum-chem. investigation reveals the complexity of mol. interactions between malvidin and polyphenols, which is preferably based on π-π and OH-π interaction moderated by repulsive Coulomb-type interactions.
- 208Kalisz, S.; Oszmiański, J.; Hładyszowski, J.; Mitek, M. Stabilization of Anthocyanin and Skullcap Flavone Complexes – Investigations with Computer Simulation and Experimental Methods Food Chem. 2013, 138 (1) 491– 500 DOI: 10.1016/j.foodchem.2012.10.146There is no corresponding record for this reference.
- 209Kunsági-Máté, S.; Ortmann, E.; Kollár, L.; Nikfardjam, M. P. Effect of the Solvatation Shell Exchange on the Formation of Malvidin- 3- O -Glucoside–Ellagic Acid Complexes J. Phys. Chem. B 2007, 111 (40) 11750– 11755 DOI: 10.1021/jp0740144There is no corresponding record for this reference.
- 210Sousa, A.; Araújo, P.; Cruz, L.; Brás, N. F.; Mateus, N.; De Freitas, V. Evidence for Copigmentation Interactions between Deoxyanthocyanidin Derivatives (Oaklins) and Common Copigments in Wine Model Solutions J. Agric. Food Chem. 2014, 62 (29) 6995– 7001 DOI: 10.1021/jf404640mThere is no corresponding record for this reference.
- 211Košinová, P.; Berka, K.; Wykes, M.; Otyepka, M.; Trouillas, P. Positioning of Antioxidant Quercetin and Its Metabolites in Lipid Bilayer Membranes: Implication for Their Lipid-Peroxidation Inhibition J. Phys. Chem. B 2012, 116 (4) 1309– 1318 DOI: 10.1021/jp208731g211Positioning of Antioxidant Quercetin and Its Metabolites in Lipid Bilayer Membranes: Implication for Their Lipid-Peroxidation InhibitionKosinova, Pavlina; Berka, Karel; Wykes, Michael; Otyepka, Michal; Trouillas, PatrickJournal of Physical Chemistry B (2012), 116 (4), 1309-1318CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Among numerous biol. activities, natural polyphenols are antioxidants widely distributed in plants capable of inhibiting lipid peroxidn., which belongs to the most serious degenerative cell processes. Positioning of antioxidants in lipid bilayers can provide an insight to the lipid-peroxidn. inhibition at the mol. level. This work aims at detg. the location and orientation of quercetin and its most representative (glucuronidated, methylated, and sulfated) metabolites in lipid bilayer via mol. dynamic simulations. We show that quercetin derivs. penetrate the lipid bilayer and that the depths of penetration depend on mol. charge and substitutional variations. In the presence of charged substituents (sulfates and glucuronidates), the mol. is pulled toward the lipid bilayer surface. The orientation also depends on substitution as H-bonds are formed between the polar head groups of the bilayer and the (i) OH groups, (ii) sugar, and (iii) sulfate moieties of the antioxidants. As flavonoids and their derivs. are preferentially localized in the lipid bilayer membrane or on the bilayer/water interface, they readily conc. in a relatively narrow membrane region. Despite the low concns. of flavonoids in food, their spatial confinement in the membrane greatly enhances their local concn. in this vital region, thus increasing their importance for in vivo biol. activities including oxidative stress defense.
- 212Sirk, T. W.; Brown, E. F.; Sum, A. K.; Friedman, M. Molecular Dynamics Study on the Biophysical Interactions of Seven Green Tea Catechins with Lipid Bilayers of Cell Membranes J. Agric. Food Chem. 2008, 56 (17) 7750– 7758 DOI: 10.1021/jf8013298212Molecular Dynamics Study on the Biophysical Interactions of Seven Green Tea Catechins with Lipid Bilayers of Cell MembranesSirk, Timothy W.; Brown, Eugene F.; Sum, Amadeu K.; Friedman, MendelJournal of Agricultural and Food Chemistry (2008), 56 (17), 7750-7758CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)Mol. dynamics simulations were performed to study the interactions of bioactive catechins (flavonoids) commonly found in green tea with lipid bilayers, as a model for cell membranes. Previously, multiple exptl. studies rationalized catechin's anticarcinogenic, antibacterial, and other beneficial effects in terms of physicochem. mol. interactions with the cell membranes. To contribute toward understanding the mol. role of catechins on the structure of cell membranes, we present simulation results for seven green tea catechins in lipid bilayer systems representative of HepG2 cancer cells. Our simulations show that the seven tea catechins evaluated have a strong affinity for the lipid bilayer via hydrogen bonding to the bilayer surface, with some of the smaller catechins able to penetrate underneath the surface. Epigallocatechin-gallate (EGCG) showed the strongest interaction with the lipid bilayer based on the no. of hydrogen bonds formed with lipid headgroups. The simulations also provide insight into the functional characteristics of the catechins that distinguish them as effective compds. to potentially alter the lipid bilayer properties. The results on the hydrogen-bonding effects, described here for the first time, may contribute to a better understanding of proposed multiple mol. mechanisms of the action of catechins in microorganisms, cancer cells, and tissues.
- 213Paloncýová, M.; Fabre, G.; DeVane, R. H.; Trouillas, P.; Berka, K.; Otyepka, M. Benchmarking of Force Fields for Molecule–Membrane Interactions J. Chem. Theory Comput. 2014, 10 (9) 4143– 4151 DOI: 10.1021/ct500419b213Benchmarking of Force Fields for Molecule-Membrane InteractionsPaloncyova, Marketa; Fabre, Gabin; DeVane, Russell H.; Trouillas, Patrick; Berka, Karel; Otyepka, MichalJournal of Chemical Theory and Computation (2014), 10 (9), 4143-4151CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Free energy profiles of eleven compds. with a model dimyristoylphosphatidylcholine (DMPC) membrane bilayer have been calcd. using five force fields, namely Berger, Slipids, CHARMM36, GAFFlipids, and GROMOS 43A1-S3 via mol. dynamics simulations. For the sake of comparison, the semicontinuous tool COSMOmic was also used. High correlation was obsd. between theor. and exptl. partition coeffs. (log K). Partition coeffs. calcd. by all-at. force fields (Slipids, CHARMM36, and GAFFlipids) and COSMOmic differed by less than 0.75 log units from the expt. and Slipids emerged as the best performing force field. This work provides the following recommendations (i) for a global, systematic and high throughput thermodn. evaluations (e.g., log K) of drugs COSMOmic is a tool of choice due to low computational costs; (ii) for studies of the hydrophilic mols. CHARMM36 should be considered; and (iii) for studies of more complex systems, taking into account all pros and cons, Slipids is the force field of choice.
- 214Podloucká, P.; Berka, K.; Fabre, G.; Paloncýová, M.; Duroux, J.-L.; Otyepka, M.; Trouillas, P. Lipid Bilayer Membrane Affinity Rationalizes Inhibition of Lipid Peroxidation by a Natural Lignan Antioxidant J. Phys. Chem. B 2013, 117 (17) 5043– 5049 DOI: 10.1021/jp3127829There is no corresponding record for this reference.
- 215Zhou, H.-X.; Gilson, M. K. Theory of Free Energy and Entropy in Noncovalent Binding Chem. Rev. 2009, 109 (9) 4092– 4107 DOI: 10.1021/cr800551w215Theory of Free Energy and Entropy in Noncovalent BindingZhou, Huan-Xiang; Gilson, Michael K.Chemical Reviews (Washington, DC, United States) (2009), 109 (9), 4092-4107CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Theory of free energy and entropy in noncovalent binding aims to support the development of well-founded models of binding and meaningful interpretation of exptl. data. Discussion includes the statistical thermodn. of binding., changes in translational and other entropy components on binding.
- 216Černỳ, J.; Hobza, P. Non-Covalent Interactions in Biomacromolecules Phys. Chem. Chem. Phys. 2007, 9 (39) 5291– 5303 DOI: 10.1039/b704781aThere is no corresponding record for this reference.
- 217Grimme, S.; Antony, J.; Schwabe, T.; Mück-Lichtenfeld, C. Density Functional Theory with Dispersion Corrections for Supramolecular Structures, Aggregates, and Complexes of (bio) Organic Molecules Org. Biomol. Chem. 2007, 5 (5) 741– 758 DOI: 10.1039/b615319bThere is no corresponding record for this reference.
- 218Sherrill, C. D. Computations of Noncovalent N Interactions Rev. Comput. Chem. 2008, 26, 1 DOI: 10.1002/9780470399545.ch1There is no corresponding record for this reference.
- 219Tschumper, G. S. Computations for Weak Noncovalent Rev. Comput. Chem. 2008, 26, 39 DOI: 10.1002/9780470399545.ch2There is no corresponding record for this reference.
- 220Foster, M. E.; Sohlberg, K. Empirically Corrected DFT and Semi-Empirical Methods for Non-Bonding Interactions Phys. Chem. Chem. Phys. 2010, 12 (2) 307– 322 DOI: 10.1039/B912859JThere is no corresponding record for this reference.
- 221Riley, K. E.; Pitoňák, M.; Jurečka, P.; Hobza, P. Stabilization and Structure Calculations for Noncovalent Interactions in Extended Molecular Systems Based on Wave Function and Density Functional Theories Chem. Rev. 2010, 110 (9) 5023– 5063 DOI: 10.1021/cr1000173221Stabilization and Structure Calculations for Noncovalent Interactions in Extended Molecular Systems Based on Wave Function and Density Functional TheoriesRiley, Kevin E.; Pitonak, Michal; Jurecka, Petr; Hobza, PavelChemical Reviews (Washington, DC, United States) (2010), 110 (9), 5023-5063CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Topics covered include: wave function theory; MP2; methods from MP3 to CCSD; d. functional theories; empirical dispersion corrections; plane wave codes and nonlocal pseudopotentials; symmetry-adapted perturbation theory; and semiempirical quantum chem. theories.
- 222Hobza, P. The Calculation of Intermolecular Interaction Energies Annu. Rep. Prog. Chem., Sect. C: Phys. Chem. 2011, 107, 148– 168 DOI: 10.1039/c1pc90005f222The calculation of intermolecular interaction energiesHobza, PavelAnnual Reports on the Progress of Chemistry, Section C: Physical Chemistry (2011), 107 (), 148-168CODEN: ACPCDW; ISSN:0260-1826. (Royal Society of Chemistry)All life on our earth can be viewed as an application of supramol. chem., with noncovalent interactions playing a central role. The knowledge of total interaction energies as well as their components is topical for understanding the nature of these interactions and, in a broader sense, also for understanding the nature of stabilization of noncovalent systems like biomacromols. Accurate data on interaction energies can only be obtained from coupled-cluster with single and double and perturbative triple excitations (CCSD(T)) calcns. performed with extended basis sets. The CCSDD(T) calcns. thus provide benchmark data which can be used for testing and/or parameterizing other, computationally economical techniques. In the present review the applicability and performance of various recently introduced wavefunction and d. functional methods are examd. in detail.
- 223Grimme, S. Density Functional Theory with London Dispersion Corrections Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1 (2) 211– 228 DOI: 10.1002/wcms.30223Density functional theory with london dispersion correctionsGrimme, StefanWiley Interdisciplinary Reviews: Computational Molecular Science (2011), 1 (2), 211-228CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Dispersion corrections to std. Kohn-Sham d. functional theory (DFT) are reviewed. The focus is on computationally efficient methods for large systems that do not depend on virtual orbitals or rely on sepd. fragments. The recommended approaches (van der Waals d. functional and DFT-D) are asymptotically correct and can be used in combination with std. or slightly modified (short-range) exchange-correlation functionals. The importance of the dispersion energy in intramol. cases (conformational problems and thermochem.) is highlighted.
- 224Klimeš, J.; Michaelides, A. Perspective: Advances and Challenges in Treating van Der Waals Dispersion Forces in Density Functional Theory J. Chem. Phys. 2012, 137 (12) 120901 DOI: 10.1063/1.4754130224Perspective: Advances and challenges in treating van der Waals dispersion forces in density functional theoryKlimes, Jiri; Michaelides, AngelosJournal of Chemical Physics (2012), 137 (12), 120901/1-120901/12CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A review. Electron dispersion forces play a crucial role in detg. the structure and properties of biomols., mol. crystals, and many other systems. However, an accurate description of dispersion is highly challenging, with the most widely used electronic structure technique, d. functional theory (DFT), failing to describe them with std. approxns. Therefore, applications of DFT to systems where dispersion is important have traditionally been of questionable accuracy. However, the last decade has seen a surge of enthusiasm in the DFT community to tackle this problem and in so-doing to extend the applicability of DFT-based methods. Here we discuss, classify, and evaluate some of the promising schemes to emerge in recent years. A brief perspective on the outstanding issues that remain to be resolved and some directions for future research are also provided. (c) 2012 American Institute of Physics.
- 225Ehrlich, S.; Moellmann, J.; Grimme, S. Dispersion-Corrected Density Functional Theory for Aromatic Interactions in Complex Systems Acc. Chem. Res. 2013, 46 (4) 916– 926 DOI: 10.1021/ar3000844225Dispersion-Corrected Density Functional Theory for Aromatic Interactions in Complex SystemsEhrlich, Stephan; Moellmann, Jonas; Grimme, StefanAccounts of Chemical Research (2013), 46 (4), 916-926CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Arom. interactions play a key role in many chem. and biol. systems. However, even if very simple models are chosen, the systems of interest are often too large to be handled with std. wave function theory (WFT). Although d. functional theory (DFT) can easily treat systems of more than 200 atoms, std. semilocal (hybrid) d. functional approxns. fail to describe the London dispersion energy, a factor that is essential for accurate predictions of inter- and intramol. noncovalent interactions. Therefore dispersion-cor. DFT provides a unique tool for the investigation and anal. of a wide range of complex arom. systems. In this Account, we start with an anal. of the noncovalent interactions in simple model dimers of hexafluorobenzene (HFB) and benzene, with a focus on electrostatic and dispersion interactions. The min. for the parallel-displaced dimers of HFB/HFB and HFB/benzene can only be explained when taking into account all contributions to the interaction energy and not by electrostatics alone. By comparison of satd. and arom. model complexes, we show that increased dispersion coeffs. for sp2-hybridized carbon atoms play a major role in arom. stacking. Modern dispersion-cor. DFT yields accurate results (about 5-10% error for the dimerization energy) for the relatively large porphyrin and coronene dimers, systems for which WFT can provide accurate ref. data only with huge computational effort. In this example, it is also demonstrated that new nonlocal, d.-dependent dispersion corrections and atom pairwise schemes mutually agree with each other. The dispersion energy is also important for the complex inter- and intramol. interactions that arise in the mol. crystals of arom. mols. In studies of hexahelicene, dispersion-cor. DFT yields "the right answer for the right reason". By comparison, std. DFT calcns. reproduce intramol. distances quite accurately in single-mol. calcns. while inter- and intramol. distances become too large when dispersion-uncorrected solid-state calcns. are carried out. Dispersion-cor. DFT can fix this problem, and these results are in excellent agreement with exptl. structure and energetic (sublimation) data. Uncorrected treatments do not even yield a bound crystal state. Finally, we present calcns. for the formation of a cationic, quadruply charged dimer of a porphyrin deriv., a case where dispersion is required in order to overcome strong electrostatic repulsion. A combination of dispersion-cor. DFT with an adequate continuum solvation model can accurately reproduce exptl. free assocn. enthalpies in soln. As in the previous examples, consideration of the electrostatic interactions alone does not provide a qual. or quant. correct picture of the interactions of this complex.
- 226DiStasio, R. A., Jr.; Gobre, V. V.; Tkatchenko, A. Many-Body van Der Waals Interactions in Molecules and Condensed Matter J. Phys.: Condens. Matter 2014, 26 (21) 213202 DOI: 10.1088/0953-8984/26/21/213202226Many-body van der Waals interactions in molecules and condensed matterDiStasio, Robert A.; Gobre, Vivekanand V.; Tkatchenko, AlexandreJournal of Physics: Condensed Matter (2014), 26 (21), 213202/1-213202/16, 16 pp.CODEN: JCOMEL; ISSN:0953-8984. (IOP Publishing Ltd.)A review. This work reviews the increasing evidence that many-body van der Waals (vdW) or dispersion interactions play a crucial role in the structure, stability and function of a wide variety of systems in biol., chem. and physics. Starting with the exact expression for the electron correlation energy provided by the adiabatic connection fluctuation-dissipation theorem, we derive both pairwise and many-body interat. methods for computing the long-range dispersion energy by considering a model system of coupled quantum harmonic oscillators within the RPA. By coupling this approach to d. functional theory, the resulting many-body dispersion (MBD) method provides an accurate and efficient scheme for computing the frequency-dependent polarizability and many-body vdW energy in mols. and materials with a finite electronic gap. A select collection of applications are presented that ascertain the fundamental importance of these non-bonded interactions across the spectrum of intermol. (the S22 and S66 benchmark databases), intramol. (conformational energies of alanine tetrapeptide) and supramol. (binding energy of the 'buckyball catcher') complexes, as well as mol. crystals (cohesive energies in oligoacenes). These applications demonstrate that electrodynamic response screening and beyond-pairwise many-body vdW interactions - both captured at the MBD level of theory - play a quant., and sometimes even qual., role in describing the properties considered herein. This work is then concluded with an in-depth discussion of the challenges that remain in the future development of reliable (accurate and efficient) methods for treating many-body vdW interactions in complex materials and provides a road-map for navigating many of the research avenues that are yet to be explored.
- 227Casimir, H.; Polder, D. The Influence of Retardation on the London-van Der Waals Forces Phys. Rev. 1948, 73 (4) 360 DOI: 10.1103/PhysRev.73.360227The influence of retardation on the London-van der Waals forcesCasimir, H. B. G.; Polder, D.Physical Review (1948), 73 (), 360-72CODEN: PHRVAO; ISSN:0031-899X.cf. C.A. 41, 1899i. Math.theoretical. The influence of retardation on the energy of interaction between two neutral atoms is investigated by means of quantum electrodynamics. In the interactions between a neutral atom and a perfectly conducting plane, and between two atoms, it is found that the influence of radiation is described by a monotonically decreasing correction factor which is equal to unity for small distances, R, compared with the wave lengths corresponding to the at. frequencies, and proportional to 1/R when R is large compared with these wave lengths.
- 228Starkschall, G.; Gordon, R. G. Calculation of Coefficients in the Power Series Expansion of the Long-Range Dispersion Force between Atoms J. Chem. Phys. 1972, 56 (6) 2801– 2806 DOI: 10.1063/1.1677610There is no corresponding record for this reference.
- 229Tang, K.; Toennies, J. P. An Improved Simple Model for the van Der Waals Potential Based on Universal Damping Functions for the Dispersion Coefficients J. Chem. Phys. 1984, 80 (8) 3726– 3741 DOI: 10.1063/1.447150229An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficientsTang, K. T.; Toennies, J. PeterJournal of Chemical Physics (1984), 80 (8), 3726-41CODEN: JCPSA6; ISSN:0021-9606.By using the earlier model (K.T.T. and J.P.T., 1977), a simple expression was derived for the radial-dependent damping functions for the individual dispersion coeffs. C2n for arbitrary even orders 2n. The damping functions are only a function of the Born-Mayer range parameter b, and thus can be applied to all systems for which this is known or can be estd. For H(1S)-H(1S), the results agree with the ab-initio damping functions of A. Koide, et al., (1981). Comparisons with less accurate previous calcns. for other systems also show agreement. By adding a Born-Mayer repulsive term [A exp(-bR)] to the damped dispersion potential, a simple universal expression was obtained for the well region of the atom-atom van der Waals potential with only 5 essential parameters A, b, C6, C8, and C10. The model was tested for the systems: H23Σ, He2, Ar2, NaK3Σ, and LiHg, for which either very precise theor. or exptl. data are available. For each system, the ab-initio dispersion coeffs. together with the parameters ε and Rm were used to det. A and b from the model potential. With these values, the reduced potentials were calcd., and found to agree with the exptl. potentials to better than 1%, and always less than the exptl. uncertainties.
- 230Thakkar, A. J.; Hettema, H.; Wormer, P. E. Abinitio Dispersion Coefficients for Interactions Involving Rare-Gas Atoms J. Chem. Phys. 1992, 97 (5) 3252– 3257 DOI: 10.1063/1.463012There is no corresponding record for this reference.
- 231Lein, M.; Dobson, J. F.; Gross, E. K. Toward the Description of van Der Waals Interactions within Density Functional Theory J. Comput. Chem. 1999, 20 (1) 12– 22 DOI: 10.1002/(SICI)1096-987X(19990115)20:1<12::AID-JCC4>3.0.CO;2-UThere is no corresponding record for this reference.
- 232Kamal, C.; Ghanty, T.; Banerjee, A.; Chakrabarti, A. The van Der Waals Coefficients between Carbon Nanostructures and Small Molecules: A Time-Dependent Density Functional Theory Study J. Chem. Phys. 2009, 131 (16) 164708 DOI: 10.1063/1.3256238There is no corresponding record for this reference.
- 233Johnson, E. R.; Becke, A. D. A Post-Hartree–Fock Model of Intermolecular Interactions J. Chem. Phys. 2005, 123 (2) 024101 DOI: 10.1063/1.1949201233A post-Hartree-Fock model of intermolecular interactionsJohnson, Erin R.; Becke, Axel D.Journal of Chemical Physics (2005), 123 (2), 024101/1-024101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Intermol. interactions are of great importance in chem. but are difficult to model accurately with computational methods. In particular, Hartree-Fock and std. d.-functional approxns. do not include the physics necessary to properly describe dispersion. These methods are sometimes cor. to account for dispersion by adding a pairwise C6/R6 term, with C6 dispersion coeffs. dependent on the atoms involved. We present a post-Hartree-Fock model in which C6 coeffs. are generated by the instantaneous dipole moment of the exchange hole. This model relies on occupied orbitals only, and involves only one, universal, empirical parameter to limit the dispersion energy at small interat. sepns. The model is extensively tested on isotropic C6 coeffs. of 178 intermol. pairs. It is also applied to the calcn. of the geometries and binding energies of 20 intermol. complexes involving dispersion, dipole-induced dipole, dipole-dipole, and hydrogen-bonding interactions, with remarkably good results.
- 234Becke, A. D.; Johnson, E. R. Exchange-Hole Dipole Moment and the Dispersion Interaction J. Chem. Phys. 2005, 122 (15) 154104 DOI: 10.1063/1.1884601234Exchange-hole dipole moment and the dispersion interactionBecke, Axel D.; Johnson, Erin R.Journal of Chemical Physics (2005), 122 (15), 154104/1-154104/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple model is presented in which the instantaneous dipole moment of the exchange hole is used to generate a dispersion interaction between nonoverlapping systems. The model is easy to implement, requiring no electron correlation (in the usual sense) or time dependence, and has been tested on various at. and mol. pairs. The resulting C6 dispersion coeffs. are remarkably accurate.
- 235Johnson, E. R. Dependence of Dispersion Coefficients on Atomic Environment J. Chem. Phys. 2011, 135 (23) 234109 DOI: 10.1063/1.3670015There is no corresponding record for this reference.
- 236Misquitta, A. J. Intermolecular Interactions. In Handbook of Computational Chemistry; Springer, 2012; pp 157– 193.There is no corresponding record for this reference.
- 237Wen, S.; Nanda, K.; Huang, Y.; Beran, G. J. Practical Quantum Mechanics-Based Fragment Methods for Predicting Molecular Crystal Properties Phys. Chem. Chem. Phys. 2012, 14 (21) 7578– 7590 DOI: 10.1039/c2cp23949c237Practical quantum mechanics-based fragment methods for predicting molecular crystal propertiesWen, Shuhao; Nanda, Kaushik; Huang, Yuanhang; Beran, Gregory J. O.Physical Chemistry Chemical Physics (2012), 14 (21), 7578-7590CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A review. Significant advances in fragment-based electronic structure methods have created a real alternative to force-field and d. functional techniques in condensed-phase problems such as mol. crystals. This perspective article highlights some of the important challenges in modeling mol. crystals and discusses techniques for addressing them. First, we survey recent developments in fragment-based methods for mol. crystals. Second, we use examples from our own recent research on a fragment-based QM/MM method, the hybrid many-body interaction (HMBI) model, to analyze the phys. requirements for a practical and effective mol. crystal model chem. We demonstrate that it is possible to predict mol. crystal lattice energies to within a couple kJ mol-1 and lattice parameters to within a few percent in small-mol. crystals. Fragment methods provide a systematically improvable approach to making predictions in the condensed phase, which is crit. to making robust predictions regarding the subtle energy differences found in mol. crystals.
- 238Wu, Q.; Yang, W. Empirical Correction to Density Functional Theory for van Der Waals Interactions J. Chem. Phys. 2002, 116 (2) 515– 524 DOI: 10.1063/1.1424928238Empirical correction to density functional theory for van der Waals interactionsWu, Qin; Yang, WeitaoJournal of Chemical Physics (2002), 116 (2), 515-524CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)An empirical method has been designed to account for the van der Waals interactions in practical mol. calcns. with d. functional theory. For each atom pair sepd. at a distance R, the method adds to the d. functional electronic structure calcns. an addnl. attraction energy EvdW = -fd(R)C6R-6, where fd(R) is the damping function which equals to one at large value of R and zero at small value of R. The coeffs. C6 for pair interactions between hydrogen, carbon, nitrogen, and oxygen atoms have been developed in this work by a least-square fitting to the mol. C6 coeffs. obtained from the dipole oscillator strength distribution method by Meath and co-workers. Two forms of the damping functions have been studied, with one dropping to zero at short distances much faster than the other. Four d. functionals have been examd.: Becke's three parameter hybrid functional with the Lee-Yang-Parr correlation functional, Becke's 1988 exchange functional with the LYP correlation functional, Becke's 1988 exchange functional with Perdew and Wang's 1991 (PW91) correlation functional, and PW91 exchange and correlation functional. The method has been applied to three systems where the van der Waals attractions are known to be important: rare-gas diat. mols., stacking of base pairs, and polyalanines' conformation stabilities. The results show that this empirical method, with the damping function dropping to zero smoothly, provides a significant correction to both of the Becke's hybrid functional and the PW91 exchange and correlation functional. Results are comparable with the corresponding second-order Moller-Plesset calcns. in many cases.
- 239Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections Phys. Chem. Chem. Phys. 2008, 10 (44) 6615– 6620 DOI: 10.1039/b810189b239Long-range corrected hybrid density functionals with damped atom-atom dispersion correctionsChai, Jeng-Da; Head-Gordon, MartinPhysical Chemistry Chemical Physics (2008), 10 (44), 6615-6620CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We report re-optimization of a recently proposed long-range cor. (LC) hybrid d. functional [J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106] to include empirical atom-atom dispersion corrections. The resulting functional, ωB97X-D yields satisfactory accuracy for thermochem., kinetics, and non-covalent interactions. Tests show that for non-covalent systems, ωB97X-D shows slight improvement over other empirical dispersion-cor. d. functionals, while for covalent systems and kinetics it performs noticeably better. Relative to our previous functionals, such as ωB97X, the new functional is significantly superior for non-bonded interactions, and very similar in performance for bonded interactions.
- 240Liu, Y.; Goddard, W. A. I. A Universal Damping Function for Empirical Dispersion Correction on Density Functional Theory Mater. Trans. 2009, 50 (7) 1664– 1670 DOI: 10.2320/matertrans.MF200911240A universal damping function for empirical dispersion correction on density functional theoryLiu, Yi; Goddard, William A., IIIMaterials Transactions (2009), 50 (7), 1664-1670CODEN: MTARCE; ISSN:1345-9678. (Japan Institute of Metals)A damped London dispersion interaction is generally adopted in empirical dispersion corrections on d. functional theory (DFT), where dispersion parameters are detd. empirically to reproduce correct dispersive interactions after assuming a damping function. The key to a successful dispersion correction is choosing an appropriate damping function. In this work we propose a single universal damping function that can represent several damping functions used in literatures with a few adjustable parameters. This universal damping function provides a unified formula that allows convenient comparison and flexible optimization in dispersion cor. DFT methods. Using the optimized universal damping functions and dispersion parameters, we develop dispersion correction methods for HF, B3LYP and PBE theories. We calc. the dispersive energies accurately for rare gas diat. mols. and benzene dimers with an averaged error <4.1%.
- 241Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu J. Chem. Phys. 2010, 132 (15) 154104 DOI: 10.1063/1.3382344241A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-PuGrimme, Stefan; Antony, Jens; Ehrlich, Stephan; Krieg, HelgeJournal of Chemical Physics (2010), 132 (15), 154104/1-154104/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The method of dispersion correction as an add-on to std. Kohn-Sham d. functional theory (DFT-D) has been refined regarding higher accuracy, broader range of applicability, and less empiricism. The main new ingredients are atom-pairwise specific dispersion coeffs. and cutoff radii that are both computed from first principles. The coeffs. for new eighth-order dispersion terms are computed using established recursion relations. System (geometry) dependent information is used for the first time in a DFT-D type approach by employing the new concept of fractional coordination nos. (CN). They are used to interpolate between dispersion coeffs. of atoms in different chem. environments. The method only requires adjustment of two global parameters for each d. functional, is asymptotically exact for a gas of weakly interacting neutral atoms, and easily allows the computation of at. forces. Three-body nonadditivity terms are considered. The method has been assessed on std. benchmark sets for inter- and intramol. noncovalent interactions with a particular emphasis on a consistent description of light and heavy element systems. The mean abs. deviations for the S22 benchmark set of noncovalent interactions for 11 std. d. functionals decrease by 15%-40% compared to the previous (already accurate) DFT-D version. Spectacular improvements are found for a tripeptide-folding model and all tested metallic systems. The rectification of the long-range behavior and the use of more accurate C6 coeffs. also lead to a much better description of large (infinite) systems as shown for graphene sheets and the adsorption of benzene on an Ag(111) surface. For graphene it is found that the inclusion of three-body terms substantially (by about 10%) weakens the interlayer binding. We propose the revised DFT-D method as a general tool for the computation of the dispersion energy in mols. and solids of any kind with DFT and related (low-cost) electronic structure methods for large systems. (c) 2010 American Institute of Physics.
- 242Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory J. Comput. Chem. 2011, 32 (7) 1456– 1465 DOI: 10.1002/jcc.21759242Effect of the damping function in dispersion corrected density functional theoryGrimme, Stefan; Ehrlich, Stephan; Goerigk, LarsJournal of Computational Chemistry (2011), 32 (7), 1456-1465CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)It is shown by an extensive benchmark on mol. energy data that the math. form of the damping function in DFT-D methods has only a minor impact on the quality of the results. For 12 different functionals, a std. "zero-damping" formula and rational damping to finite values for small interat. distances according to Becke and Johnson (BJ-damping) has been tested. The same (DFT-D3) scheme for the computation of the dispersion coeffs. is used. The BJ-damping requires one fit parameter more for each functional (three instead of two) but has the advantage of avoiding repulsive interat. forces at shorter distances. With BJ-damping better results for nonbonded distances and more clear effects of intramol. dispersion in four representative mol. structures are found. For the noncovalently-bonded structures in the S22 set, both schemes lead to very similar intermol. distances. For noncovalent interaction energies BJ-damping performs slightly better but both variants can be recommended in general. The exception to this is Hartree-Fock that can be recommended only in the BJ-variant and which is then close to the accuracy of cor. GGAs for non-covalent interactions. According to the thermodn. benchmarks BJ-damping is more accurate esp. for medium-range electron correlation problems and only small and practically insignificant double-counting effects are obsd. It seems to provide a phys. correct short-range behavior of correlation/dispersion even with unmodified std. functionals. In any case, the differences between the two methods are much smaller than the overall dispersion effect and often also smaller than the influence of the underlying d. functional. © 2011 Wiley Periodicals, Inc.; J. Comput. Chem., 2011.
- 243Johnson, E. R.; Becke, A. D. A Post-Hartree-Fock Model of Intermolecular Interactions: Inclusion of Higher-Order Corrections J. Chem. Phys. 2006, 124 (17) 174104 DOI: 10.1063/1.2190220243A post-Hartree-Fock model of intermolecular interactions: Inclusion of higher-order correctionsJohnson, Erin R.; Becke, Axel D.Journal of Chemical Physics (2006), 124 (17), 174104/1-174104/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We have previously demonstrated that the dipole moment of the exchange hole can be used to derive intermol. C6 dispersion coeffs. [J. Chem. Phys. 122, 154104 (2005)]. This was subsequently the basis for a novel post-Hartree-Fock model of intermol. interactions [J. Chem. Phys. 123, 024101 (2005)]. In the present work, the model is extended to include higher-order dispersion coeffs. C8 and C10. The extended model performs very well for prediction of intermonomer sepns. and binding energies of 45 van der Waals complexes. In particular, it performs twice as well as basis-set extrapolated MP2 theory for dispersion-bound complexes, with minimal computational cost.
- 244Grimme, S. Accurate Description of van Der Waals Complexes by Density Functional Theory Including Empirical Corrections J. Comput. Chem. 2004, 25 (12) 1463– 1473 DOI: 10.1002/jcc.20078244Accurate description of van der Waals complexes by density functional theory including empirical correctionsGrimme, StefanJournal of Computational Chemistry (2004), 25 (12), 1463-1473CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)An empirical method to account for van der Waals interactions in practical calcns. in the framework of the d. functional theory (termed DFT-D) was tested for a wide variety of mol. complexes. As in previous schemes, the dispersive energy was described by damped interat. potentials of the form C6R-6. The use of pure, gradient-cor. d. functionals (BLYP and PBE), together with the resoln.-of-the-identity (RI) approxn. for the Coulomb operator, allows very efficient computations for large systems. In contrast to the previous work, extended AO basis sets of polarized TZV or QZV quality were employed, which reduced the basis set superposition error to a negligible extend. By using a global scaling factor for the at. C6 coeffs., the functional dependence of the results could be strongly reduced. The "double counting" of correlation effects for strongly bound complexes was found to be insignificant if steep damping functions were employed. The method was applied to a total of 29 complexes of atoms and small mols. (Ne, CH4, NH3, H2O, CH3F, N2, F2, formic acid, ethene, and ethine) with each other and with benzene, to benzene, naphthalene, pyrene, and coronene dimers, the naphthalene trimer, coronene·H2O and four H-bonded and stacked DNA base pairs (AT and GC). In almost all cases, very good agreement with reliable theor. or exptl. results for binding energies and intermol. distances is obtained. For stacked arom. systems and the important base pairs, the DFT-D-BLYP model seems to be even superior to std. MP2 treatments that systematically over-bind. The good results obtained suggest the approach as a practical tool to describe the properties of many important van der Waals systems in chem. Furthermore, the DFT-D data may either be used to calibrate much simpler (e.g., force-field) potentials or the optimized structures can be used as input for more accurate ab initio calcns. of the interaction energies.
- 245Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction J. Comput. Chem. 2006, 27 (15) 1787– 1799 DOI: 10.1002/jcc.20495245Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 246Jurečka, P.; Černỳ, J.; Hobza, P.; Salahub, D. R. Density Functional Theory Augmented with an Empirical Dispersion Term. Interaction Energies and Geometries of 80 Noncovalent Complexes Compared with Ab Initio Quantum Mechanics Calculations J. Comput. Chem. 2007, 28 (2) 555– 569 DOI: 10.1002/jcc.20570There is no corresponding record for this reference.
- 247Tkatchenko, A.; Scheffler, M. Accurate Molecular van Der Waals Interactions from Ground-State Electron Density and Free-Atom Reference Data Phys. Rev. Lett. 2009, 102 (7) 073005 DOI: 10.1103/PhysRevLett.102.073005247Accurate Molecular Van Der Waals Interactions from Ground-State Electron Density and Free-Atom Reference DataTkatchenko, Alexandre; Scheffler, MatthiasPhysical Review Letters (2009), 102 (7), 073005/1-073005/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We present a parameter-free method for an accurate detn. of long-range van der Waals interactions from mean-field electronic structure calcns. Our method relies on the summation of interat. C6 coeffs., derived from the electron d. of a mol. or solid and accurate ref. data for the free atoms. The mean abs. error in the C6 coeffs. is 5.5% when compared to accurate exptl. values for 1225 intermol. pairs, irresp. of the employed exchange-correlation functional. We show that the effective at. C6 coeffs. depend strongly on the bonding environment of an atom in a mol. Finally, we analyze the van der Waals radii and the damping function in the C6R-6 correction method for d.-functional theory calcns.
- 248Steinmann, S. N.; Corminboeuf, C. A System-Dependent Density-Based Dispersion Correction J. Chem. Theory Comput. 2010, 6 (7) 1990– 2001 DOI: 10.1021/ct1001494248A System-Dependent Density-Based Dispersion CorrectionSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Theory and Computation (2010), 6 (7), 1990-2001CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)D. functional approxns. fail to provide a consistent description of weak mol. interactions arising from small electron d. overlaps. A simple remedy to correct for the missing interactions is to add a posteriori an attractive energy term summed over all atom pairs in the system. The d.-dependent energy correction, presented herein, is applicable to all elements of the periodic table and is easily combined with any electronic structure method, which lacks the accurate treatment of weak interactions. Dispersion coeffs. are computed according to Becke and Johnson's exchange-hole dipole moment (XDM) formalism, thereby depending on the chem. environment of an atom (d., oxidn. state). The long-range ∼R-6 potential is supplemented with higher-order correction terms (∼R-8 and ∼R-10) through the universal damping function of Tang and Toennies. A genuine damping factor depending on (iterative) Hirshfeld (overlap) populations, at. ionization energies, and two adjustable parameters specifically fitted to a given DFT functional is also introduced. The proposed correction, dDXDM, dramatically improves the performance of popular d. functionals. The anal. of 30 (dispersion cor.) d. functionals on 145 systems reveals that dDXDM largely reduces the errors of the parent functionals for both inter- and intramol. interactions. With mean abs. deviations (MADs) of 0.74-0.84 kcal mol-1, PBE-dDXDM, PBE0-dDXDM, and B3LYP-dDXDM outperform the computationally more demanding and most recent functionals such as M06-2X and B2PLYP-D (MAD of 1.93 and 1.06 kcal mol-1, resp.).
- 249Steinmann, S. N.; Corminboeuf, C. Comprehensive Benchmarking of a Density-Dependent Dispersion Correction J. Chem. Theory Comput. 2011, 7 (11) 3567– 3577 DOI: 10.1021/ct200602x249Comprehensive Benchmarking of a Density-Dependent Dispersion CorrectionSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Theory and Computation (2011), 7 (11), 3567-3577CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Std. d. functional approxns. cannot accurately describe interactions between nonoverlapping densities. A simple remedy consists in correcting for the missing interactions a posteriori, adding an attractive energy term summed over all atom pairs. The d.-dependent energy correction, dDsC, presented herein, is constructed from dispersion coeffs. computed on the basis of a generalized gradient approxn. to Becke and Johnson's exchange-hole dipole moment formalism. DDsC also relies on an extended Tang and Toennies damping function accounting for charge-overlap effects. The comprehensive benchmarking on 341 diverse reaction energies divided into 18 illustrative test sets validates the robust performance and general accuracy of dDsC for describing various intra- and intermol. interactions. With a total MAD of 1.3 kcal mol-1, B97-dDsC slightly improves the results of M06-2X and B2PLYP-D3 (MAD = 1.4 kcal mol-1 for both) at a lower computational cost. The d. dependence of both the dispersion coeffs. and the damping function makes the approach esp. valuable for modeling redox reactions and charged species in general.
- 250Steinmann, S. N.; Corminboeuf, C. A Generalized-Gradient Approximation Exchange Hole Model for Dispersion Coefficients J. Chem. Phys. 2011, 134 (4) 044117 DOI: 10.1063/1.3545985250A generalized-gradient approximation exchange hole model for dispersion coefficientsSteinmann, Stephan N.; Corminboeuf, ClemenceJournal of Chemical Physics (2011), 134 (4), 044117/1-044117/5CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A simple method for computing accurate d.-dependent dispersion coeffs. is presented. The dispersion coeffs. are modeled by a generalized gradient-type approxn. to Becke and Johnson's exchange hole dipole moment formalism. Our most cost-effective variant, based on a disjoint description of atoms in a mol., gives mean abs. errors in the C6 coeffs. for 90 complexes below 10%. The inclusion of the missing long-range van der Waals interactions in d. functionals using the derived coeffs. in a pair wise correction leads to highly accurate typical noncovalent interaction energies. (c) 2011 American Institute of Physics.
- 251Liu, Y.; Goddard, W. A., III First-Principles-Based Dispersion Augmented Density Functional Theory: From Molecules to Crystals J. Phys. Chem. Lett. 2010, 1 (17) 2550– 2555 DOI: 10.1021/jz100615g251First-Principles-Based Dispersion Augmented Density Functional Theory: From Molecules to CrystalsLiu, Yi; Goddard, William A., IIIJournal of Physical Chemistry Letters (2010), 1 (17), 2550-2555CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Std. implementations of d. functional theory (DFT) describe well strongly bound mols. and solids but fail to describe long-range van der Waals attractions. We propose here first-principles-based augmentation to DFT that leads to the proper long-range 1/R6 attraction of the London dispersion while leading to low gradients (small forces) at normal valence distances so that it preserves the accurate geometries and thermochem. of std. DFT methods. The DFT-low gradient (DFT-lg) formula differs from previous DFT-D methods by using a purely attractive dispersion correction while not affecting valence bond distances. We demonstrate here that the DFT-lg model leads to good descriptions for graphite, benzene, naphthalene, and anthracene crystals, using just three parameters fitted to reproduce the full potential curves of high-level ab initio quantum mechanics [CCSD(T)] on gas-phase benzene dimers. The addnl. computational costs for this DFT-lg formalism are negligible.
- 252Kim, H.; Choi, J.-M.; Goddard, W. A., III Universal Correction of Density Functional Theory to Include London Dispersion (up to Lr, Element 103) J. Phys. Chem. Lett. 2012, 3 (3) 360– 363 DOI: 10.1021/jz2016395252Universal Correction of Density Functional Theory to Include London Dispersion (up to Lr, Element 103)Kim, Hyungjun; Choi, Jeong-Mo; Goddard, William A., IIIJournal of Physical Chemistry Letters (2012), 3 (3), 360-363CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Conventional d. functional theory (DFT) fails to describe accurately the London dispersion essential for describing mol. interactions in soft matter (biol. systems, polymers, nucleic acids) and mol. crystals. This has led to several methods in which atom-dependent potentials are added into the Kohn-Sham DFT energy. Some of these corrections were fitted to accurate quantum mech. results, but it will be tedious to det. the appropriate parameters to describe all of the atoms of the periodic table. We propose an alternative approach in which a single parameter in the low-gradient (lg) functional form is combined with the rule-based UFF (universal force-field) nonbond parameters developed for the entire periodic table (up to Lr, Z = 103), named as a DFT-ulg method. We show that DFT-ulg method leads to a very accurate description of the properties for mol. complexes and mol. crystals, providing the means for predicting more accurate weak interactions across the periodic table.
- 253Román-Pérez, G.; Soler, J. M. Efficient Implementation of a van Der Waals Density Functional: Application to Double-Wall Carbon Nanotubes Phys. Rev. Lett. 2009, 103 (9) 096102 DOI: 10.1103/PhysRevLett.103.096102253Efficient Implementation of a van der Waals Density Functional: Application to Double-Wall Carbon NanotubesRoman-Perez, Guillermo; Soler, Jose M.Physical Review Letters (2009), 103 (9), 096102/1-096102/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We present an efficient implementation of the van der Waals d. functional of Dion et al., which expresses the nonlocal correlation energy as a double spatial integral. We factorize the integration kernel and use fast Fourier transforms to evaluate the self-consistent potential, total energy, and at. forces, in O(NlogN) operations. The resulting overhead, for medium and large systems, is a small fraction of the total computational cost, representing a dramatic speedup over the O(N2) evaluation of the double integral. This opens the realm of first-principles simulations to the large systems of interest in soft matter and biomol. problems. We apply the method to calc. the binding energies and the barriers for relative translation and rotation in double-wall carbon nanotubes.
- 254Andersson, Y.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Interactions in Density-Functional Theory Phys. Rev. Lett. 1996, 76 (1) 102– 105 DOI: 10.1103/PhysRevLett.76.102There is no corresponding record for this reference.
- 255Dobson, J. F.; Dinte, B. P. Constraint Satisfaction in Local and Gradient Susceptibility Approximations: Application to a van Der Waals Density Functional Phys. Rev. Lett. 1996, 76 (11) 1780 DOI: 10.1103/PhysRevLett.76.1780There is no corresponding record for this reference.
- 256Sato, T.; Tsuneda, T.; Hirao, K. Van Der Waals Interactions Studied by Density Functional Theory Mol. Phys. 2005, 103 (6–8) 1151– 1164 DOI: 10.1080/00268970412331333474There is no corresponding record for this reference.
- 257Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Density Functional for General Geometries Phys. Rev. Lett. 2004, 92 (24) 246401 DOI: 10.1103/PhysRevLett.92.246401257Van der Waals Density Functional for General GeometriesDion, M.; Rydberg, H.; Schroeder, E.; Langreth, D. C.; Lundqvist, B. I.Physical Review Letters (2004), 92 (24), 246401/1-246401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)A scheme within d. functional theory is proposed that provides a practical way to generalize to unrestricted geometries the method applied with some success to layered geometries [H. Rydberg et al., Phys. Rev. Lett. 91, 126402 (2003)]. It includes van der Waals forces in a seamless fashion. By expansion to second order in a carefully chosen quantity contained in the long-range part of the correlation functional, the nonlocal correlations are expressed in terms of a d.-d. interaction formula. It contains a relatively simple parametrized kernel, with parameters detd. by the local d. and its gradient. The proposed functional is applied to rare gas and benzene dimers, where it is shown to give a realistic description.
- 258Pernal, K.; Podeszwa, R.; Patkowski, K.; Szalewicz, K. Dispersionless Density Functional Theory Phys. Rev. Lett. 2009, 103 (26) 263201 DOI: 10.1103/PhysRevLett.103.263201There is no corresponding record for this reference.
- 259Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van Der Waals Density Functional J. Phys.: Condens. Matter 2010, 22 (2) 022201 DOI: 10.1088/0953-8984/22/2/022201259Chemical accuracy for the van der Waals density functionalKlimes, Jiri; Bowler, David R.; Michaelides, AngelosJournal of Physics: Condensed Matter (2010), 22 (2), 022201/1-022201/5CODEN: JCOMEL; ISSN:0953-8984. (Institute of Physics Publishing)The non-local van der Waals d. functional (vdW-DF) of Dion et al is a very promising scheme for the efficient treatment of dispersion bonded systems. We show here that the accuracy of vdW-DF can be dramatically improved both for dispersion and hydrogen bonded complexes through the judicious selection of its underlying exchange functional. New and published exchange functionals are identified that deliver much better than chem. accuracy from vdW-DF for the S22 benchmark set of weakly interacting dimers and for water clusters. Improved performance for the adsorption of water on salt is also obtained.
- 260Lee, K.; Murray, É. D.; Kong, L.; Lundqvist, B. I.; Langreth, D. C. Higher-Accuracy van Der Waals Density Functional Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 82 (8) 081101 DOI: 10.1103/PhysRevB.82.081101260Higher-accuracy van der Waals density functionalLee, Kyuho; Murray, Eamonn D.; Kong, Lingzhu; Lundqvist, Bengt I.; Langreth, David C.Physical Review B: Condensed Matter and Materials Physics (2010), 82 (8), 081101/1-081101/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)We propose a second version of the van der Waals d. functional of Dion et al. [Phys. Rev. Lett. 92, 246401 (2004)], employing a more accurate semilocal exchange functional and the use of a large-N asymptote gradient correction in detg. the vdW kernel. The predicted binding energy, equil. sepn., and potential-energy curve shape are close to those of accurate quantum chem. calcns. on 22 duplexes. We anticipate the enabling of chem. accurate calcns. in sparse materials of importance for condensed matter, surface, chem., and biol. physics.
- 261Hamada, I. Van Der Waals Density Functional Made Accurate Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 89 (12) 121103 DOI: 10.1103/PhysRevB.89.121103261van der Waals density functional made accurateHamada, IkutaroPhysical Review B: Condensed Matter and Materials Physics (2014), 89 (12), 121103/1-121103/5CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)I propose a van der Waals d. functional (vdW-DF) that improves upon the description of energetics and geometries of mols., solids, and adsorption systems over the original vdW-DF. The functional is based on the nonlocal correlation for the second version of the vdW-DF and an exchange functional that recovers the second-order gradient expansion approxn. in the slowly varying limit, while reproducing the large d. gradient behavior proposed by Becke. A systematic assessment of the proposed functional is presented, which demonstrates the applicability of the proposed vdW-DF to a wide range of systems.
- 262Vydrov, O. A.; Van Voorhis, T. Nonlocal van Der Waals Density Functional Made Simple Phys. Rev. Lett. 2009, 103 (6) 063004 DOI: 10.1103/PhysRevLett.103.063004262Nonlocal van der Waals Density Functional Made SimpleVydrov, Oleg A.; Van Voorhis, TroyPhysical Review Letters (2009), 103 (6), 063004/1-063004/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We derive a nonlocal correlation functional that adequately describes van der Waals interactions not only in the asymptotic long-range regime but also at short range. Unlike its precursor, developed by Langreth, Lundqvist, and co-workers, the new functional has a simple analytic form, finite for all interelectron sepns., well behaved in the slowly varying d. limit, and generalized to spin-polarized systems.
- 263Vydrov, O. A.; Van Voorhis, T. Improving the Accuracy of the Nonlocal van Der Waals Density Functional with Minimal Empiricism J. Chem. Phys. 2009, 130 (10) 104105 DOI: 10.1063/1.3079684There is no corresponding record for this reference.
- 264Vydrov, O. A.; Van Voorhis, T. Implementation and Assessment of a Simple Nonlocal van Der Waals Density Functional J. Chem. Phys. 2010, 132 (16) 164113 DOI: 10.1063/1.3398840There is no corresponding record for this reference.
- 265Vydrov, O. A.; Van Voorhis, T. Nonlocal van Der Waals Density Functional: The Simpler the Better J. Chem. Phys. 2010, 133 (24) 244103 DOI: 10.1063/1.3521275265Nonlocal van der Waals density functional: The simpler the betterVydrov, Oleg A.; Van Voorhis, TroyJournal of Chemical Physics (2010), 133 (24), 244103/1-244103/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We devise a nonlocal correlation energy functional that describes the entire range of dispersion interactions in a seamless fashion using only the electron d. as input. The new functional is considerably simpler than its predecessors of a similar type. The functional has a tractable and robust analytic form that lends itself to efficient self-consistent implementation. When paired with an appropriate exchange functional, our nonlocal correlation model yields accurate interaction energies of weakly-bound complexes, not only near the energy min. but also far from equil. Our model exhibits an outstanding precision at predicting equil. intermonomer sepns. in van der Waals complexes. It also gives accurate covalent bond lengths and atomization energies. Hence the functional proposed in this work is a computationally inexpensive electronic structure tool of broad applicability. (c) 2010 American Institute of Physics.
- 266Sabatini, R.; Gorni, T.; de Gironcoli, S. Nonlocal van Der Waals Density Functional Made Simple and Efficient Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87 (4) 041108 DOI: 10.1103/PhysRevB.87.041108266Nonlocal van der Waals density functional made simple and efficientSabatini, Riccardo; Gorni, Tommaso; de Gironcoli, StefanoPhysical Review B: Condensed Matter and Materials Physics (2013), 87 (4), 041108/1-041108/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)We present a simple revision of the VV10 nonlocal d. functional by Vydrov and Van Voorhis for dispersion interactions. Unlike the original functional our modification allows nonlocal correlation energy and its derivs. to be efficiently evaluated in a plane wave framework along the lines pioneered by Roman-Perez and Soler. Our revised functional maintains the outstanding precision of the original VV10 in noncovalently bound complexes and performs well in representative covalent, ionic, and metallic solids.
- 267Hujo, W.; Grimme, S. Performance of the van Der Waals Density Functional VV10 and (hybrid) GGA Variants for Thermochemistry and Noncovalent Interactions J. Chem. Theory Comput. 2011, 7 (12) 3866– 3871 DOI: 10.1021/ct200644w267Performance of the van der Waals Density Functional VV10 and (hybrid)GGA Variants for Thermochemistry and Noncovalent InteractionsHujo, Waldemar; Grimme, StefanJournal of Chemical Theory and Computation (2011), 7 (12), 3866-3871CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal van der Waals d. functional VV10 (Vydrov, O. A.; Van Voorhis, T. J. Chem. Phys.2010, 133, 244103) is tested for the thermochem. properties of 1200 + atoms and mols. in the GMTKN30 database in order to assess its global accuracy. Five GGA and hybrid functionals in unmodified form are augmented by the nonlocal (NL) part of the VV10 functional (one parameter adjusted). The addn. of the NL dispersion energy definitely improves the results of all tested functionals. On the basis of little empiricism and basic phys. insight, DFT-NL can be recommended as a fully electronic, robust electronic structure method.
- 268Aragó, J.; Ortí, E.; Sancho-García, J. C. Nonlocal van Der Waals Approach Merged with Double-Hybrid Density Functionals: Toward the Accurate Treatment of Noncovalent Interactions J. Chem. Theory Comput. 2013, 9 (8) 3437– 3443 DOI: 10.1021/ct4003527268Nonlocal van der Waals Approach Merged with Double-Hybrid Density Functionals: Toward the Accurate Treatment of Noncovalent InteractionsArago, Juan; Orti, Enrique; Sancho-Garcia, Juan C.Journal of Chemical Theory and Computation (2013), 9 (8), 3437-3443CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Noncovalent interactions drive the self-assembly of weakly interacting mol. systems to form supramol. aggregates, which play a major role in nanotechnol. and biochem. In this work, we present a thorough assessment of the performance of different double-hybrid d. functionals (PBE0-DH-NL, revPBE0-DH-NL, B2PLYP-NL, and TPSS0-DH-NL), as well as their parent hybrid and (meta)GGA functionals, in combination with the most modern version of the nonlocal (NL) van der Waals correction. It is shown that this nonlocal correction can be successfully coupled with double-hybrid d. functionals thanks to the short-range attenuation parameter b, which has been optimized against ref. interaction energies of benchmarking mol. complexes (S22 and S66 databases). Among all the double-hybrid functionals evaluated, revPBE0-DH-NL and B2PLYP-NL behave remarkably accurate with mean unsigned errors (MUE) as small as 0.20 kcal/mol for the training sets and in the 0.25-0.42 kcal/mol range for an independent database (NCCE31). They can be thus seen as appropriate functionals to use in a broad no. of applications where noncovalent interactions play an important role. Overall, the nonlocal van der Waals approach combined with last-generation d. functionals is confirmed as an accurate and affordable computational tool for the modeling of weakly bonded mol. systems.
- 269Cooper, V. R. Van Der Waals Density Functional: An Appropriate Exchange Functional Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81 (16) 161104 DOI: 10.1103/PhysRevB.81.161104269Van der Waals density functional: An appropriate exchange functionalCooper, Valentino R.Physical Review B: Condensed Matter and Materials Physics (2010), 81 (16), 161104/1-161104/4CODEN: PRBMDO; ISSN:1098-0121. (American Physical Society)In this Rapid Communication, an exchange functional which is compatible with the nonlocal Rutgers-Chalmers correlation functional [van der Waals d. functional (vdW-DF)] is presented. This functional, when employed with vdW-DF, demonstrates remarkable improvements on intermol. sepn. distances while further improving the accuracy of vdW-DF interaction energies. The key to the success of this three-parameter functional is its redn. in short-range exchange repulsion through matching to the gradient expansion approxn. in the slowly varying/high-d. limit while recovering the large reduced gradient, s, limit set in the revised Perdew-Burke-Ernzerhof (revPBE) exchange functional. This augmented exchange functional could be a soln. to long-standing issues of vdW-DF lending to further applicability of d.-functional theory to the study of relatively large, dispersion bound (van der Waals) complexes.
- 270Vydrov, O. A.; Van Voorhis, T. Benchmark Assessment of the Accuracy of Several van Der Waals Density Functionals J. Chem. Theory Comput. 2012, 8 (6) 1929– 1934 DOI: 10.1021/ct300081y270Benchmark Assessment of the Accuracy of Several van der Waals Density FunctionalsVydrov, Oleg A.; Van Voorhis, TroyJournal of Chemical Theory and Computation (2012), 8 (6), 1929-1934CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal correlation functional VV10, developed recently in our group, describes the whole range of dispersion interactions in a seamless and general fashion using only the electron d. as input. The VV10 functional has a simple analytic form that can be adjusted for pairing with the exchange functional of choice. In this paper, we use several benchmark data sets of weakly interacting mol. complexes to test the accuracy of two VV10 variants, differing in their treatment of the exchange component. For the sake of comparison, several other d. functionals suitable for noncovalent interactions were also tested against the same benchmarks. We find that the "default" version of VV10 with semilocal exchange gives very accurate geometries and binding energies for most van der Waals complexes but systematically overbinds hydrogen-bonded complexes. The alternative variant of VV10 with long-range cor. hybrid exchange performs exceptionally well for all types of weak bonding sampled in this study, including hydrogen bonds.
- 271Hujo, W.; Grimme, S. Performance of Non-Local and Atom-Pairwise Dispersion Corrections to DFT for Structural Parameters of Molecules with Noncovalent Interactions J. Chem. Theory Comput. 2013, 9 (1) 308– 315 DOI: 10.1021/ct300813c271Performance of Non-Local and Atom-Pairwise Dispersion Corrections to DFT for Structural Parameters of Molecules with Noncovalent InteractionsHujo, Waldemar; Grimme, StefanJournal of Chemical Theory and Computation (2013), 9 (1), 308-315CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The nonlocal, electron d. dependent dispersion correction of Vydrov and Van Voorhis (Vydrov, O. A.; Van Voorhis, T. J. Chem. Phys.2010, 133, 244103), termed VV10 or DFT-NL, was implemented for structural optimizations of mols. It was tested in combination with the four (hybrid)GGA d. functionals TPSS, TPSS0, B3LYP, and revPBE38 for inter- and intramol. noncovalent interactions (NCI) and compared to results from atom-pairwise dispersion cor. DFT-D3. The methods were applied to a wide range of different problems, namely the S22 and S66 test sets, large transition metal complexes, water hexamer clusters, hexahelicene, and four other difficult cases of intramol. NCI. Crit. interat. distances were computed remarkably accurately by both dispersion corrections compared to theor. or exptl. ref. data and inter- and intramol. interactions were treated on equal footing. The methods can be recommended as reliable and robust tools for geometry optimizations of large systems in which long-range dispersion forces are crucial.
- 272Tran, F.; Hutter, J. Nonlocal van Der Waals Functionals: The Case of Rare-Gas Dimers and Solids J. Chem. Phys. 2013, 138 (20) 204103 DOI: 10.1063/1.4807332There is no corresponding record for this reference.
- 273Gobre, V. V.; Tkatchenko, A. Scaling Laws for van Der Waals Interactions in Nanostructured Materials Nat. Commun. 2013, 4, 2341 DOI: 10.1038/ncomms3341273Scaling laws for van der Waals interactions in nanostructured materialsGobre Vivekanand V; Tkatchenko AlexandreNature communications (2013), 4 (), 2341 ISSN:.Van der Waals interactions have a fundamental role in biology, physics and chemistry, in particular in the self-assembly and the ensuing function of nanostructured materials. Here we utilize an efficient microscopic method to demonstrate that van der Waals interactions in nanomaterials act at distances greater than typically assumed, and can be characterized by different scaling laws depending on the dimensionality and size of the system. Specifically, we study the behaviour of van der Waals interactions in single-layer and multilayer graphene, fullerenes of varying size, single-wall carbon nanotubes and graphene nanoribbons. As a function of nanostructure size, the van der Waals coefficients follow unusual trends for all of the considered systems, and deviate significantly from the conventionally employed pairwise-additive picture. We propose that the peculiar van der Waals interactions in nanostructured materials could be exploited to control their self-assembly.
- 274Ambrosetti, A.; Alfè, D.; DiStasio, R. A., Jr.; Tkatchenko, A. Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular Systems J. Phys. Chem. Lett. 2014, 5 (5) 849– 855 DOI: 10.1021/jz402663k274Hard Numbers for Large Molecules: Toward Exact Energetics for Supramolecular SystemsAmbrosetti, Alberto; Alfe, Dario; DiStasio, Robert A.; Tkatchenko, AlexandreJournal of Physical Chemistry Letters (2014), 5 (5), 849-855CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Noncovalent interactions are ubiquitous in mol. and condensed-phase environments, and hence a reliable theor. description of these fundamental interactions could pave the way toward a more complete understanding of the microscopic underpinnings for a diverse set of systems in chem. and biol. In this work, we demonstrate that recent algorithmic advances coupled to the availability of large-scale computational resources make the stochastic quantum Monte Carlo approach to solving the Schrodinger equation an optimal contender for attaining "chem. accuracy" (1 kcal/mol) in the binding energies of supramol. complexes of chem. relevance. To illustrate this point, we considered a select set of seven host-guest complexes, representing the spectrum of noncovalent interactions, including dispersion or van der Waals forces, π-π stacking, hydrogen bonding, hydrophobic interactions, and electrostatic (ion-dipole) attraction. A detailed anal. of the interaction energies reveals that a complete theor. description necessitates treatment of terms well beyond the std. London and Axilrod-Teller contributions to the van der Waals dispersion energy.
- 275Šimová, L.; Řezáč, J.; Hobza, P. Convergence of the Interaction Energies in Noncovalent Complexes in the Coupled-Cluster Methods Up to Full Configuration Interaction J. Chem. Theory Comput. 2013, 9 (8) 3420– 3428 DOI: 10.1021/ct4002762275Convergence of the Interaction Energies in Noncovalent Complexes in the Coupled-Cluster Methods Up to Full Configuration InteractionSimova, Lucia; Rezac, Jan; Hobza, PavelJournal of Chemical Theory and Computation (2013), 9 (8), 3420-3428CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The CCSD(T) method stands out among various coupled-cluster (CC) approxns. as the golden std. in computational chem. and is widely and successfully used in the realm of covalent and noncovalent interactions. The CCSD(T) method provides reliable interaction energies, but their surprising accuracy is believed to arise partially from an error compensation. The convergence of the CC expansion has been investigated up to fully iterative pentuple excitations (CCSDTQP); for the smallest eight electron complexes, the full CI calcns. have also been performed. We conclude that the convergence of interaction energy at noncovalent accuracy (0.01 kcal/mol) for the complexes studied is reached already at CCSDTQ or CCSDT(Q) levels. When even higher accuracy (spectroscopic accuracy of 1 cm-1, or 3 cal/mol) is required, then the noniterative CCSDTQ(P) method could be used.
- 276Řezáč, J.; Šimová, L.; Hobza, P. CCSD[T] Describes Noncovalent Interactions Better than the CCSD(T), CCSD(TQ), and CCSDT Methods J. Chem. Theory Comput. 2013, 9 (1) 364– 369 DOI: 10.1021/ct3008777276CCSD[T] Describes Noncovalent Interactions Better than the CCSD(T), CCSD(TQ), and CCSDT MethodsRezac, Jan; Simova, Lucia; Hobza, PavelJournal of Chemical Theory and Computation (2013), 9 (1), 364-369CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The CCSD(T) method is often called the "gold std." of computational chem., because it is one of the most accurate methods applicable to reasonably large mols. It is particularly useful for the description of noncovalent interactions where the inclusion of triple excitations is necessary for achieving a satisfactory accuracy. While it is widely used as a benchmark, the accuracy of CCSD(T) interaction energies has not been reliably quantified yet against more accurate calcns. In this work, we compare the CCSD[T], CCSD(T), and CCSD(TQ) noniterative methods with full CCSDTQ and CCSDT(Q) calcns. We investigate various types of noncovalent complexes [hydrogen-bonded (water dimer, ammonia dimer, water ··· ammonia), dispersion-bound (methane dimer, methane ··· ammonia), and π-π stacked (ethene dimer)] using various coupled-clusters schemes up to CCSDTQ in 6-31G*(0.25), 6-31G**(0.25, 0.15), and aug-cc-pVDZ basis sets. We show that CCSDT(Q) reproduces the CCSDTQ results almost exactly and can thus serve as a benchmark in the cases where CCSDTQ calcns. are not feasible. Surprisingly, the CCSD[T] method provides better agreement with the benchmark values than the other noniterative analogs, CCSD(T) and CCSD(TQ), and even than the much more expensive iterative CCSDT scheme. The CCSD[T] interaction energies differ from the benchmark data by less than 5 cal/mol on av. (for all complexes and all basis sets), whereas the error of CCSD(T) is 9 cal/mol. In larger systems, the difference between these two methods can grow by as much as 0.15 kcal/mol. While this effect can be explained only as an error compensation, the CCSD[T] method certainly deserves more attention in accurate calcns. of noncovalent interactions.
- 277Kong, L.; Bischoff, F. A.; Valeev, E. F. Explicitly Correlated R12/F12 Methods for Electronic Structure Chem. Rev. 2012, 112 (1) 75– 107 DOI: 10.1021/cr200204r277Explicitly Correlated R12/F12 Methods for Electronic StructureKong, Liguo; Bischoff, Florian A.; Valeev, Edward F.Chemical Reviews (Washington, DC, United States) (2012), 112 (1), 75-107CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The following topics are discussed: Electron correlation in atoms and mols.(CI wave functions and their symmetries); Explicitly correlated methods and wave functions; Core technol. of modern R12 methods; Coupled-Cluster R12 methods; Multireference R12 methods (R12-MRCI, MRMP2-F12, [2]R12, CASPT2-F12 and MRCI-F12, G-CASSCF, MR CABS Single correction, comparison of MR-R12 Methods).
- 278Burns, L. A.; Marshall, M. S.; Sherrill, C. D. Appointing Silver and Bronze Standards for Noncovalent Interactions: A Comparison of Spin-Component-Scaled (SCS), Explicitly Correlated (F12), and Specialized Wavefunction Approaches J. Chem. Phys. 2014, 141 (23) 234111 DOI: 10.1063/1.4903765278Appointing silver and bronze standards for noncovalent interactions: A comparison of spin-component-scaled (SCS), explicitly correlated (F12), and specialized wavefunction approachesBurns, Lori A.; Marshall, Michael S.; Sherrill, C. DavidJournal of Chemical Physics (2014), 141 (23), 234111/1-234111/21CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A systematic examn. of noncovalent interactions as modeled by wavefunction theory is presented in comparison to gold-std. quality benchmarks available for 345 interaction energies of 49 bimol. complexes. Quantum chem. techniques examd. include spin-component-scaling (SCS) variations on second-order perturbation theory (MP2) [SCS, SCS(N), SCS(MI)] and coupled cluster singles and doubles (CCSD) [SCS, SCS(MI)]; also, method combinations designed to improve dispersion contacts [DW-MP2, MP2C, MP2.5, DW-CCSD(T)-F12]; where available, explicitly correlated (F12) counterparts are also considered. Dunning basis sets augmented by diffuse functions are employed for all accessible ζ-levels; truncations of the diffuse space are also considered. After examn. of both accuracy and performance for 394 model chemistries, SCS(MI)-MP2/cc-pVQZ can be recommended for general use, having good accuracy at low cost and no ill-effects such as imbalance between hydrogen-bonding and dispersion-dominated systems or non-parallelity across dissocn. curves. Moreover, when benchmarking accuracy is desirable but gold-std. computations are unaffordable, this work recommends silver-std. [DW-CCSD(T**)-F12/aug-cc-pVDZ] and bronze-std. [MP2C-F12/aug-cc-pVDZ] model chemistries, which support accuracies of 0.05 and 0.16 kcal/mol and efficiencies of 97.3 and 5.5 h for adenine·thymine, resp. Choice comparisons of wavefunction results with the best symmetry-adapted perturbation theory [T. M. Parker, L. A. Burns, R. M. Parrish, A. G. Ryno, and C. D. Sherrill, J. Chem. Phys.140, 094106 (2014)] and d. functional theory [L. A. Burns, A. Vazquez-Mayagoitia, B. G. Sumpter, and C. D. Sherrill, J. Chem. Phys.134, 084107 (2011)] methods previously studied for these databases are provided for readers' guidance. (c) 2014 American Institute of Physics.
- 279Austin, B. M.; Zubarev, D. Y.; Lester, W. A., Jr. Quantum Monte Carlo and Related Approaches Chem. Rev. 2012, 112 (1) 263– 288 DOI: 10.1021/cr2001564279Quantum Monte Carlo and Related ApproachesAustin, Brian M.; Zubarev, Dmitry Yu.; Lester, William A., Jr.Chemical Reviews (Washington, DC, United States) (2012), 112 (1), 263-288CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The following topics are discussed: Variational Monte Carlo, Fixed-node diffusion Monte Carlo, Self-healing diffusion Monte Carlo, Auxiliary field quantum Monte Carlo, Reptation quantum Monte Carlo, Full CI quantum Monte Carlo, Time-dependent quantum Monte Carlo; Trial electronic wave functions (antisym., backflow transformed, Jastrow), Trial wave function optimization, Effective core potentials; Computational considerations (Linear scaling quantum Monte Carlo, Parallelization and hardware acceleration, Advances in algorithms and software); Applications.
- 280Dubeckỳ, M.; Jurečka, P.; Derian, R.; Hobza, P.; Otyepka, M.; Mitas, L. Quantum Monte Carlo Methods Describe Noncovalent Interactions with Subchemical Accuracy J. Chem. Theory Comput. 2013, 9 (10) 4287– 4292 DOI: 10.1021/ct4006739280Quantum Monte Carlo Methods Describe Noncovalent Interactions with Subchemical AccuracyDubecky, Matus; Jurecka, Petr; Derian, Rene; Hobza, Pavel; Otyepka, Michal; Mitas, LubosJournal of Chemical Theory and Computation (2013), 9 (10), 4287-4292CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An accurate description of noncovalent interaction energies is one of the most challenging tasks in computational chem. To date, nonempirical CCSD-(T)/CBS has been used as a benchmark ref. However, its practical use is limited due to the rapid growth of its computational cost with the system complexity. Here, we show that the fixed-node diffusion Monte Carlo (FN-DMC) method with a more favorable scaling is capable of reaching the CCSD-(T)/CBS within subchem. accuracy (<0.1 kcal/mol) on a testing set of six small noncovalent complexes including the water dimer. In benzene/water, benzene/methane, and the T-shape benzene dimer, FN-DMC provides interaction energies that agree within 0.25 kcal/mol with the best available CCSD-(T)/CBS ests. The demonstrated predictive power of FN-DMC therefore provides new opportunities for studies of the vast and important class of medium/large noncovalent complexes.
- 281Dubeckỳ, M.; Derian, R.; Jurečka, P.; Mitas, L.; Hobza, P.; Otyepka, M. Quantum Monte Carlo for Noncovalent Interactions: An Efficient Protocol Attaining Benchmark Accuracy Phys. Chem. Chem. Phys. 2014, 16 (38) 20915– 20923 DOI: 10.1039/C4CP02093FThere is no corresponding record for this reference.
- 282Neese, F.; Wennmohs, F.; Hansen, A. Efficient and Accurate Local Approximations to Coupled-Electron Pair Approaches: An Attempt to Revive the Pair Natural Orbital Method J. Chem. Phys. 2009, 130 (11) 114108 DOI: 10.1063/1.3086717282Efficient and accurate local approximations to coupled-electron pair approaches: An attempt to revive the pair natural orbital methodNeese, Frank; Wennmohs, Frank; Hansen, AndreasJournal of Chemical Physics (2009), 130 (11), 114108/1-114108/18CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Coupled-electron pair approxns. (CEPAs) and coupled-pair functionals (CPFs) have been popular in the 1970s and 1980s and have yielded excellent results for small mols. Recently, interest in CEPA and CPF methods has been renewed. It has been shown that these methods lead to competitive thermochem., kinetic, and structural predictions. They greatly surpass second order Moller-Plesset and popular d. functional theory based approaches in accuracy and are intermediate in quality between CCSD and CCSD(T) in extended benchmark studies. In this work an efficient prodn. level implementation of the closed shell CEPA and CPF methods is reported that can be applied to medium sized mols. in the range of 50-100 atoms and up to about 2000 basis functions. The internal space is spanned by localized internal orbitals. The external space is greatly compressed through the method of pair natural orbitals (PNOs) that was also introduced by the pioneers of the CEPA approaches. Our implementation also makes extended use of d. fitting (or resoln. of the identity) techniques in order to speed up the laborious integral transformations. The method is called local pair natural orbital CEPA (LPNO-CEPA) (LPNO-CPF). The implementation is centered around the concepts of electron pairs and matrix operations. Altogether three cutoff parameters are introduced that control the size of the significant pair list, the av. no. of PNOs per electron pair, and the no. of contributing basis functions per PNO. With the conservatively chosen default values of these thresholds, the method recovers about 99.8% of the canonical correlation energy. This translates to abs. deviations from the canonical result of only a few kcal mol-1. Extended numerical test calcns. demonstrate that LPNO-CEPA (LPNO-CPF) has essentially the same accuracy as parent CEPA (CPF) methods for thermochem., kinetics, weak interactions, and potential energy surfaces but is up to 500 times faster. The method performs best in conjunction with large and flexible basis sets. These results open the way for large-scale chem. applications. (c) 2009 American Institute of Physics.
- 283Neese, F.; Hansen, A.; Liakos, D. G. Efficient and Accurate Approximations to the Local Coupled Cluster Singles Doubles Method Using a Truncated Pair Natural Orbital Basis J. Chem. Phys. 2009, 131 (6) 064103 DOI: 10.1063/1.3173827283Efficient and accurate approximations to the local coupled cluster singles doubles method using a truncated pair natural orbital basisNeese, Frank; Hansen, Andreas; Liakos, Dimitrios G.Journal of Chemical Physics (2009), 131 (6), 064103/1-064103/15CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A prodn. level implementation of the closed-shell local quadratic CI and coupled cluster methods with single and double excitations (QCISD and CCSD) based on the concept of pair natural orbitals (local pair natural orbital LPNO-QCISD and LPNO-CCSD) is reported, evaluated, and discussed. This work is an extension of the earlier developed LPNO coupled-electron pair approxn. (LNPO-CEPA) method and makes extended use of the resoln. of the identity (RI) or d. fitting (DF) approxn. Two variants of each method are compared. The less accurate approxns. (LPNO2-QCISD/LPNO2-CCSD) still recover 98.7%-99.3% of the correlation energy in the given basis and have modest disk space requirements. The more accurate variants (LPNO1-QCISD/LPNO1-CCSD) typically recover 99.75%-99.95% of the correlation energy in the given basis but require the Coulomb and exchange operators with up to two-external indexes to be stored on disk. Both variants have comparable computational efficiency. The convergence of the results with respect to the natural orbital truncation parameter (TCutPNO) has been studied. Extended numerical tests have been performed on abs. and relative correlation energies as function of basis set size and TCutPNO as well as on reaction energies, isomerization energies, and weak intermol. interactions. The results indicate that the errors of the LPNO methods compared to the canonical QCISD and CCSD methods are below 1 kcal/mol with our default thresholds. Finally, some calcns. on larger mols. are reported (ranging from 40-86 atoms) and it is shown that for medium sized mols. the total wall clock time required to complete the LPNO-CCSD calcns. is only two to four times that of the preceding SCF. Thus these methods are highly suitable for large-scale computational chem. applications. Since there are only three thresholds involved that have been given conservative default values, the methods can be confidentially used in a "black-box" fashion in the same way as their canonical counterparts. (c) 2009 American Institute of Physics.
- 284Huntington, L. M.; Nooijen, M. pCCSD: Parameterized Coupled-Cluster Theory with Single and Double Excitations J. Chem. Phys. 2010, 133 (18) 184109 DOI: 10.1063/1.3494113There is no corresponding record for this reference.
- 285Huntington, L. M. J.; Hansen, A.; Neese, F.; Nooijen, M. Accurate Thermochemistry from a Parameterized Coupled-Cluster Singles and Doubles Model and a Local Pair Natural Orbital Based Implementation for Applications to Larger Systems J. Chem. Phys. 2012, 136 (6) 064101 DOI: 10.1063/1.3682325There is no corresponding record for this reference.
- 286Hansen, A.; Liakos, D. G.; Neese, F. Efficient and Accurate Local Single Reference Correlation Methods for High-Spin Open-Shell Molecules Using Pair Natural Orbitals J. Chem. Phys. 2011, 135 (21) 214102 DOI: 10.1063/1.3663855286Efficient and accurate local single reference correlation methods for high-spin open-shell molecules using pair natural orbitalsHansen, Andreas; Liakos, Dimitrios G.; Neese, FrankJournal of Chemical Physics (2011), 135 (21), 214102/1-214102/20CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A prodn. level implementation of the high-spin open-shell (spin unrestricted) single ref. coupled pair, quadratic CI and coupled cluster methods with up to doubly excited determinants in the framework of the local pair natural orbital (LPNO) concept is reported. This work is an extension of the closed-shell LPNO methods developed earlier. The internal space is spanned by localized orbitals, while the external space for each electron pair is represented by a truncated PNO expansion. The laborious integral transformation assocd. with the large no. of PNOs becomes feasible through the extensive use of d. fitting (resoln. of the identity (RI)) techniques. Tech. complications arising for the open-shell case and the use of quasi-restricted orbitals for the construction of the ref. determinant are discussed in detail. As in the closed-shell case, only three cutoff parameters control the av. no. of PNOs per electron pair, the size of the significant pair list, and the no. of contributing auxiliary basis functions per PNO. The chosen threshold default values ensure robustness and the results of the parent canonical methods are reproduced to high accuracy. Comprehensive numerical tests on abs. and relative energies as well as timings consistently show that the outstanding performance of the LPNO methods carries over to the open-shell case with minor modifications. Finally, hyperfine couplings calcd. with the variational LPNO-CEPA/1 method, for which a well-defined expectation value type d. exists, indicate the great potential of the LPNO approach for the efficient calcn. of mol. properties. (c) 2011 American Institute of Physics.
- 287Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural Triple Excitations in Local Coupled Cluster Calculations with Pair Natural Orbitals J. Chem. Phys. 2013, 139 (13) 134101 DOI: 10.1063/1.4821834287Natural triple excitations in local coupled cluster calculations with pair natural orbitalsRiplinger, Christoph; Sandhoefer, Barbara; Hansen, Andreas; Neese, FrankJournal of Chemical Physics (2013), 139 (13), 134101/1-134101/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In this work, the extension of the previously developed domain based local pair-natural orbital (DLPNO) based singles- and doubles coupled cluster (DLPNO-CCSD) method to perturbatively include connected triple excitations is reported. The development is based on the concept of triples-natural orbitals that span the joint space of the three pair natural orbital (PNO) spaces of the three electron pairs that are involved in the calcn. of a given triple-excitation contribution. The truncation error is very smooth and can be significantly reduced through extrapolation to the zero threshold. However, the extrapolation procedure does not improve relative energies. The overall computational effort of the method is asymptotically linear with the system size O(N). Actual linear scaling has been confirmed in test calcns. on alkane chains. The accuracy of the DLPNO-CCSD(T) approxn. relative to semicanonical CCSD(T0) is comparable to the previously developed DLPNO-CCSD method relative to canonical CCSD. Relative energies are predicted with an av. error of approx. 0.5 kcal/mol for a challenging test set of medium sized org. mols. The triples correction typically adds 30%-50% to the overall computation time. Thus, very large systems can be treated on the basis of the current implementation. In addn. to the linear C150H302 (452 atoms, >8800 basis functions) we demonstrate the first CCSD(T) level calcn. on an entire protein, Crambin with 644 atoms, and more than 6400 basis functions.
- 288Sparta, M.; Neese, F. Chemical Applications Carried out by Local Pair Natural Orbital Based Coupled-Cluster Methods Chem. Soc. Rev. 2014, 43 (14) 5032– 5041 DOI: 10.1039/c4cs00050aThere is no corresponding record for this reference.
- 289Liakos, D. G.; Hansen, A.; Neese, F. Weak Molecular Interactions Studied with Parallel Implementations of the Local Pair Natural Orbital Coupled Pair and Coupled Cluster Methods J. Chem. Theory Comput. 2011, 7 (1) 76– 87 DOI: 10.1021/ct100445s289Weak molecular interactions studied with parallel implementations of the local pair natural orbital coupled pair and coupled cluster methodsLiakos, Dimitrios G.; Hansen, Andreas; Neese, FrankJournal of Chemical Theory and Computation (2011), 7 (1), 76-87CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A parallel implementation of the recently developed local pair natural orbital CEPA (LPNO-CEPA/n, n = Version 1, 2, or 3) and the corresponding LPNO coupled cluster method with single- and double excitations (LPNO-CCSD) is described. A detailed anal. alongside pseudocode is presented for the most important computational steps. The scaling with respect to the no. of processors is reasonable and speedups of about 10 with 14 processors have been found in benchmark calcns. (wall-clock time). The most important factor limiting the efficiency of the scaling with respect to the no. of processors is probably the limited bandwidth of the presently prevailing multicore machines. The parallel LPNO methods were applied to study weak intermol. interactions. Initially, the well-established S22 set of mols. was studied. The mean abs. error resulting from the use of the LPNO-CEPA/1 method relative to the most recent CCSD(T) ref. data is found to be 0.24 kcal/mol. Thus, LPNO-CEPA/1 holds great promise for the efficient ab initio treatment of weak intermol. interactions. In order to demonstrate the applicability of the methods to real systems, a two-dimensional potential energy surface for a trimer of 2,4-dihydroxy-3-acetyl-6-Me acetophenone [C11H12O4] (81 atoms, 1296 basis functions, 133 single points) has been calcd. with the LPNO-CEPA/1 method. In this system, a clear distinction can be made between hydrogen bonding and π-π interactions. The global min. on the PES obtained from the calcns. agrees excellently with the exptl. detd. crystal structure. By contrast, popular d. functional methods show no discernible min.
- 290Liakos, D. G.; Neese, F. Improved Correlation Energy Extrapolation Schemes Based on Local Pair Natural Orbital Methods J. Phys. Chem. A 2012, 116 (19) 4801– 4816 DOI: 10.1021/jp302096v290Improved Correlation Energy Extrapolation Schemes Based on Local Pair Natural Orbital MethodsLiakos, Dimitrios G.; Neese, FrankJournal of Physical Chemistry A (2012), 116 (19), 4801-4816CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)It is well-known that the basis set limit is difficult to reach in correlated post Hartree-Fock ab initio calcns. One possible route forward is to employ basis set extrapolation schemes. In order to avoid prohibitively expensive calcns., the highest level calcn. (typically based on the "gold std." coupled cluster theory with single, double, and perturbative triple excitations, CCSD(T)) is only performed with the smallest basis set, and the remaining basis set incompleteness is estd. at a lower level of theory, typically second-order Moeller-Plesset perturbation theory (MP2). In this work, we provide a comprehensive investigation of alternative schemes where the MP2 extrapolation is replaced by the coupled-electron pair approxn., version 1 (CEPA/1) or the local pair natural orbital version of this method (LPNO-CEPA/1). It is shown that the MP2 method achieves apparent accuracy only due to error cancellation. Systematically more accurate results at small addnl. computational cost are obtained if the MP2 step is replaced by LPNO-CEPA/1. The errors of LPNO-CEPA/1 relative to canonical CEPA/1 are negligible. Owing to the highly systematic nature of the deviations between canonical and LPNO methods, basis set extrapolation reduces the LPNO errors in the total energies by 1 order of magnitude (∼0.2 kcal/mol) and errors in energy differences to essentially zero. Using the CCSD(T)/LPNO-CEPA/1-based extrapolation scheme, new ref. values are proposed for the recently published S66 set of interaction energies. The deviations between the new values and the original interactions energies are mostly very small but reach values up to 0.3 kcal/mol.
- 291Schwabe, T. Accurate and Fast Treatment of Large Molecular Systems: Assessment of CEPA and pCCSD within the Local Pair Natural Orbital Approximation J. Comput. Chem. 2012, 33 (26) 2067– 2072 DOI: 10.1002/jcc.23042There is no corresponding record for this reference.
- 292Riplinger, C.; Neese, F. An Efficient and near Linear Scaling Pair Natural Orbital Based Local Coupled Cluster Method J. Chem. Phys. 2013, 138 (3) 034106 DOI: 10.1063/1.4773581292An efficient and near linear scaling pair natural orbital based local coupled cluster methodRiplinger, Christoph; Neese, FrankJournal of Chemical Physics (2013), 138 (3), 034106/1-034106/18CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In previous publications, it was shown that an efficient local coupled cluster method with single- and double excitations can be based on the concept of pair natural orbitals (PNOs) . The resulting local pair natural orbital-coupled-cluster single double (LPNO-CCSD) method has since been proven to be highly reliable and efficient. For large mols., the no. of amplitudes to be detd. is reduced by a factor of 105-106 relative to a canonical CCSD calcn. on the same system with the same basis set. In the original method, the PNOs were expanded in the set of canonical virtual orbitals and single excitations were not truncated. This led to a no. of fifth order scaling steps that eventually rendered the method computationally expensive for large mols. (e.g., >100 atoms). In the present work, these limitations are overcome by a complete redesign of the LPNO-CCSD method. The new method is based on the combination of the concepts of PNOs and projected AOs (PAOs). Thus, each PNO is expanded in a set of PAOs that in turn belong to a given electron pair specific domain. In this way, it is possible to fully exploit locality while maintaining the extremely high compactness of the original LPNO-CCSD wavefunction. No terms are dropped from the CCSD equations and domains are chosen conservatively. The correlation energy loss due to the domains remains below <0.05%, which implies typically 15-20 but occasionally up to 30 atoms per domain on av. The new method has been given the acronym DLPNO-CCSD ("domain based LPNO-CCSD"). The method is nearly linear scaling with respect to system size. The original LPNO-CCSD method had three adjustable truncation thresholds that were chosen conservatively and do not need to be changed for actual applications. In the present treatment, no addnl. truncation parameters have been introduced. Any addnl. truncation is performed on the basis of the three original thresholds. There are no real-space cutoffs. Single excitations are truncated using singles-specific natural orbitals. Pairs are prescreened according to a multipole expansion of a pair correlation energy est. based on local orbital specific virtual orbitals (LOSVs). Like its LPNO-CCSD predecessor, the method is completely of black box character and does not require any user adjustments. It is shown here that DLPNO-CCSD is as accurate as LPNO-CCSD while leading to computational savings exceeding one order of magnitude for larger systems. The largest calcns. reported here featured >8800 basis functions and >450 atoms. In all larger test calcns. done so far, the LPNO-CCSD step took less time than the preceding Hartree-Fock calcn., provided no approxns. have been introduced in the latter. Thus, based on the present development reliable CCSD calcns. on large mols. with unprecedented efficiency and accuracy are realized. (c) 2013 American Institute of Physics.
- 293Sancho-García, J. C.; Aragó, J.; Ortí, E.; Olivier, Y. Obtaining the Lattice Energy of the Anthracene Crystal by Modern yet Affordable First-Principles Methods J. Chem. Phys. 2013, 138 (20) 204304 DOI: 10.1063/1.4806436There is no corresponding record for this reference.
- 294Grimme, S. Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory Chem. - Eur. J. 2012, 18 (32) 9955– 9964 DOI: 10.1002/chem.201200497294Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional TheoryGrimme, StefanChemistry - A European Journal (2012), 18 (32), 9955-9964, S9955/1-S9955/53CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The equil. assocn. free enthalpies ΔGa for typical supramol. complexes in soln. are calcd. by ab initio quantum chem. methods. Ten neutral and three pos. charged complexes with exptl. ΔGa values in the range 0 to -21 kcal mol-1 (on av. -6 kcal mol-1) are investigated. The theor. approach employs a (non-dynamic) single-structure model, but computes the various energy terms accurately without any special empirical adjustments. Dispersion cor. d. functional theory (DFT-D3) with extended basis sets (triple-ζ and quadruple-ζ quality) is used to det. structures and gas-phase interaction energies (ΔE), the COSMO-RS continuum solvation model (based on DFT data) provides solvation free enthalpies and the remaining ro-vibrational enthalpic/entropic contributions are obtained from harmonic frequency calcns. Low-lying vibrational modes are treated by a free-rotor approxn. The accurate account of London dispersion interactions is mandatory with contributions in the range -5 to -60 kcal mol-1 (up to 200% of ΔE). Inclusion of three-body dispersion effects improves the results considerably. A semi-local (TPSS) and a hybrid d. functional (PW6B95) have been tested. Although the ΔGa values result as a sum of individually large terms with opposite sign (ΔE vs. solvation and entropy change), the approach provides unprecedented accuracy for ΔGa values with errors of only 2 kcal mol-1 on av. Relative affinities for different guests inside the same host are always obtained correctly. The procedure is suggested as a predictive tool in supramol. chem. and can be applied routinely to semirigid systems with 300-400 atoms. The various contributions to binding and enthalpy-entropy compensations are discussed.
- 295Calbo, J.; Ortí, E.; Sancho-García, J. C.; Aragó, J. Accurate Treatment of Large Supramolecular Complexes by Double-Hybrid Density Functionals Coupled with Nonlocal van Der Waals Corrections J. Chem. Theory Comput. 2015, 11 (3) 932– 939 DOI: 10.1021/acs.jctc.5b00002295Accurate Treatment of Large Supramolecular Complexes by Double-Hybrid Density Functionals Coupled with Nonlocal van der Waals CorrectionsCalbo, Joaquin; Orti, Enrique; Sancho-Garcia, Juan C.; Arago, JuanJournal of Chemical Theory and Computation (2015), 11 (3), 932-939CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present a thorough assessment of the performance of some representative double-hybrid d. functionals (revPBE0-DH-NL and B2PLYP-NL) as well as their parent hybrid and GGA counterparts, in combination with the most modern version of the nonlocal (NL) van der Waals correction to describe very large weakly interacting mol. systems dominated by noncovalent interactions. Prior to the assessment, an accurate and homogeneous set of ref. interaction energies was computed for the supramol. complexes constituting the L7 and S12L data sets by using the novel, precise, and efficient DLPNO-CCSD(T) method at the complete basis set limit (CBS). The correction of the basis set superposition error and the inclusion of the deformation energies (for the S12L set) have been crucial for obtaining precise DLPNO-CCSD(T)/CBS interaction energies. Among the d. functionals evaluated, the double-hybrid revPBE0-DH-NL and B2PLYP-NL with the three-body dispersion correction provide remarkably accurate assocn. energies very close to the chem. accuracy. Overall, the NL van der Waals approach combined with proper d. functionals can be seen as an accurate and affordable computational tool for the modeling of large weakly bonded supramol. systems.
- 296Pitoňák, M.; Řezáč, J.; Hobza, P. Spin-Component Scaled Coupled-Clusters Singles and Doubles Optimized towards Calculation of Noncovalent Interactions Phys. Chem. Chem. Phys. 2010, 12 (33) 9611– 9614 DOI: 10.1039/c0cp00158aThere is no corresponding record for this reference.
- 297Grimme, S. Improved Second-Order Møller–Plesset Perturbation Theory by Separate Scaling of Parallel-and Antiparallel-Spin Pair Correlation Energies J. Chem. Phys. 2003, 118 (20) 9095– 9102 DOI: 10.1063/1.1569242There is no corresponding record for this reference.
- 298Szabados, Á. Theoretical Interpretation of Grimme’s Spin-Component-Scaled Second Order Møller-Plesset Theory J. Chem. Phys. 2006, 125 (21) 214105 DOI: 10.1063/1.2404660298Theoretical interpretation of Grimme's spin-component-scaled second order Moller-Plesset theorySzabados, AgnesJournal of Chemical Physics (2006), 125 (21), 214105/1-214105/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)It is shown that spin-component-scaled second order Moller-Plesset theory proposed by Grimme [J. Chem. Phys. 118, 9095 (2003)] can be interpreted as a two-parameter scaling of the zero order Hamiltonian, a generalization of the approach reported by Feenberg [Phys. Rev. 103, 1116 (1956)].
- 299Fink, R. F. Spin-Component-Scaled Møller–Plesset (SCS-MP) Perturbation Theory: A Generalization of the MP Approach with Improved Properties J. Chem. Phys. 2010, 133 (17) 174113 DOI: 10.1063/1.3503041There is no corresponding record for this reference.
- 300Hill, J. G.; Platts, J. A. Spin-Component Scaling Methods for Weak and Stacking Interactions J. Chem. Theory Comput. 2007, 3 (1) 80– 85 DOI: 10.1021/ct6002737300Spin-Component Scaling Methods for Weak and Stacking InteractionsHill, J. Grant; Platts, James A.Journal of Chemical Theory and Computation (2007), 3 (1), 80-85CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)New scaling parameters are presented for use in the spin-component scaled (SCS) variant of d. fitted local second-order Moller-Plesset perturbation theory (DF-LMP2) that have been optimized for use in evaluating the interaction energy between nucleic acid base pairs. The optimal set of parameters completely neglects the contribution from antiparallel-spin electron pairs to the MP2 energy while scaling the parallel contribution by 1.76. These spin-component scaled for nucleobases (SCSN) parameters are obtained by minimizing, with respect to SCS parameters, the rms interaction energy error relative to the best available literature values, over a set of ten stacked nucleic acid base pairs. The applicability of this scaling to a wide variety of noncovalent interactions is verified through evaluation of a larger set of model complexes, including those dominated by dispersion and electrostatics.
- 301Hill, J. G.; Platts, J. A. Calculating Stacking Interactions in Nucleic Acid Base-Pair Steps Using Spin-Component Scaling and Local Second Order Møller–Plesset Perturbation Theory Phys. Chem. Chem. Phys. 2008, 10 (19) 2785– 2791 DOI: 10.1039/b718691fThere is no corresponding record for this reference.
- 302Karton, A.; Tarnopolsky, A.; Lamere, J.-F.; Schatz, G. C.; Martin, J. M. Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical Kinetics J. Phys. Chem. A 2008, 112 (50) 12868– 12886 DOI: 10.1021/jp801805p302Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical KineticsKarton, Amir; Tarnopolsky, Alex; Lamere, Jean-Francois; Schatz, George C.; Martin, Jan M. L.Journal of Physical Chemistry A (2008), 112 (50), 12868-12886CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We present a no. of near-exact, nonrelativistic, Born-Oppenheimer ref. data sets for the parametrization of more approx. methods (such as DFT functionals). The data were obtained by means of the W4 ab initio computational thermochem. protocol, which has a 95% confidence interval well below 1 kJ/mol. Our data sets include W4-08, which are total atomization energies of over 100 small mols. that cover varying degrees of non-dynamical correlations, and DBH24-W4, which are W4 theory values for Truhlar's set of 24 representative barrier heights. The usual procedure of comparing calcd. DFT values with exptl. atomization energies is hampered by comparatively large exptl. uncertainties in many exptl. values and compds. errors due to deficiencies in the DFT functional with those resulting from neglect of relativity and finite nuclear mass. Comparison with accurate, explicitly nonrelativistic, ab initio data avoids these issues. We then proceed to explore the performance of B2x-PLYP-type double hybrid functionals for atomization energies and barrier heights. The optimum hybrids for hydrogen-transfer reactions, heavy-atoms transfers, nucleophilic substitutions, and unimol. and recombination reactions are quite different from one another: out of these subsets, the heavy-atom transfer reactions are by far the most sensitive to the percentages of Hartree-Fock-type exchange y and MP2-type correlation x in an (x,y) double hybrid. The (42,72) hybrid B2K-PLYP, as reported in a preliminary communication, represents the best compromise between thermochem. and hydrogen-transfer barriers, while also yielding excellent performance for nucleophilic substitutions. By optimizing for best overall performance on both thermochem. and the DBH24-W4 data set, however, we find a new (36,65) hybrid which we term B2GP-PLYP. At a slight expense in performance for hydrogen-transfer barrier heights and nucleophilic substitutions, we obtain substantially better performance for the other reaction types. Although both B2K-PLYP and B2GP-PLYP are capable of 2 kcal/mol quality thermochem., B2GP-PLYP appears to be the more robust toward non-dynamical correlation and strongly polar character. We addnl. find that double-hybrid functionals display excellent performance for such problems as hydrogen bonding, prototype late transition metal reactions, pericyclic reactions, prototype cumulene-polyacetylene system, and weak interactions.
- 303Distasio, R. A., Jr.; Head-Gordon, M. Optimized Spin-Component Scaled Second-Order Møller-Plesset Perturbation Theory for Intermolecular Interaction Energies Mol. Phys. 2007, 105 (8) 1073– 1083 DOI: 10.1080/00268970701283781There is no corresponding record for this reference.
- 304Marchetti, O.; Werner, H.-J. Accurate Calculations of Intermolecular Interaction Energies Using Explicitly Correlated Coupled Cluster Wave Functions and a Dispersion-Weighted MP2 Method J. Phys. Chem. A 2009, 113 (43) 11580– 11585 DOI: 10.1021/jp9059467There is no corresponding record for this reference.
- 305Riley, K. E.; Platts, J. A.; Řezáč, J.; Hobza, P.; Hill, J. G. Assessment of the Performance of MP2 and MP2 Variants for the Treatment of Noncovalent Interactions J. Phys. Chem. A 2012, 116 (16) 4159– 4169 DOI: 10.1021/jp211997bThere is no corresponding record for this reference.
- 306Takatani, T.; Hohenstein, E. G.; Sherrill, C. D. Improvement of the Coupled-Cluster Singles and Doubles Method via Scaling Same-and Opposite-Spin Components of the Double Excitation Correlation Energy J. Chem. Phys. 2008, 128 (12) 124111 DOI: 10.1063/1.2883974There is no corresponding record for this reference.
- 307Jung, Y.; Lochan, R. C.; Dutoi, A. D.; Head-Gordon, M. Scaled Opposite-Spin Second Order Møller–Plesset Correlation Energy: An Economical Electronic Structure Method J. Chem. Phys. 2004, 121 (20) 9793– 9802 DOI: 10.1063/1.1809602There is no corresponding record for this reference.
- 308Lochan, R. C.; Jung, Y.; Head-Gordon, M. Scaled Opposite Spin Second Order Møller-Plesset Theory with Improved Physical Description of Long-Range Dispersion Interactions J. Phys. Chem. A 2005, 109 (33) 7598– 7605 DOI: 10.1021/jp0514426There is no corresponding record for this reference.
- 309Takatani, T.; Sherrill, C. D. Performance of Spin-Component-Scaled Møller–Plesset Theory (SCS-MP2) for Potential Energy Curves of Noncovalent Interactions Phys. Chem. Chem. Phys. 2007, 9 (46) 6106– 6114 DOI: 10.1039/b709669kThere is no corresponding record for this reference.
- 310Antony, J.; Grimme, S. Is Spin-Component Scaled Second-Order Møller-Plesset Perturbation Theory an Appropriate Method for the Study of Noncovalent Interactions in Molecules? J. Phys. Chem. A 2007, 111 (22) 4862– 4868 DOI: 10.1021/jp070589pThere is no corresponding record for this reference.
- 311Bachorz, R. A.; Bischoff, F. A.; Höfener, S.; Klopper, W.; Ottiger, P.; Leist, R.; Frey, J. A.; Leutwyler, S. Scope and Limitations of the SCS-MP2 Method for Stacking and Hydrogen Bonding Interactions Phys. Chem. Chem. Phys. 2008, 10 (19) 2758– 2766 DOI: 10.1039/b718494hThere is no corresponding record for this reference.
- 312King, R. A. On the Accuracy of Spin-Component-Scaled Perturbation Theory (SCS-MP2) for the Potential Energy Surface of the Ethylene Dimer Mol. Phys. 2009, 107 (8–12) 789– 795 DOI: 10.1080/00268970802641242There is no corresponding record for this reference.
- 313Grabowski, I.; Fabiano, E.; Sala, F. D. A Simple Non-Empirical Procedure for Spin-Component-Scaled MP2 Methods Applied to the Calculation of the Dissociation Energy Curve of Noncovalently-Interacting Systems Phys. Chem. Chem. Phys. 2013, 15 (37) 15485– 15493 DOI: 10.1039/c3cp51431eThere is no corresponding record for this reference.
- 314Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van Der Waals Complexes Chem. Rev. 1994, 94 (7) 1887– 1930 DOI: 10.1021/cr00031a008314Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van der Waals ComplexesJeziorski, Bogumil; Moszynski, Robert; Szalewicz, KrzysztofChemical Reviews (Washington, DC, United States) (1994), 94 (7), 1887-930CODEN: CHREAY; ISSN:0009-2665.The topics reviewed with 445 refs. include: polarization theory; exchange effects; multipole expansion of the interaction energy; charge-overlap effects and bipolar expansion of polarization energies; the intramonomer electron correlation problem and many-body formulation of symmetry-adapted perturbation theory; and applications.
- 315Misquitta, A. J.; Podeszwa, R.; Jeziorski, B.; Szalewicz, K. Intermolecular Potentials Based on Symmetry-Adapted Perturbation Theory with Dispersion Energies from Time-Dependent Density-Functional Calculations J. Chem. Phys. 2005, 123 (21) 214103 DOI: 10.1063/1.2135288315Intermolecular potentials based on symmetry-adapted perturbation theory with dispersion energies from time-dependent density-functional calculationsMisquitta, Alston J.; Podeszwa, Rafal; Jeziorski, Bogumil; Szalewicz, KrzysztofJournal of Chemical Physics (2005), 123 (21), 214103/1-214103/14CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Recently, three of us have proposed a method [Phys. Rev. Lett. 91, 33201 (2003)] for an accurate calcn. of the dispersion energy utilizing frequency-dependent d. susceptibilities of monomers obtained from time-dependent d.-functional theory (DFT). In the present paper, we report numerical calcns. for the helium, neon, water, and carbon dioxide dimers and show that for a wide range of intermonomer sepns., including the van der Waals and short-range repulsion regions, the method provides dispersion energies with accuracies comparable to those that can be achieved using the current most sophisticated wave-function methods. If the dispersion energy is combined with (i) the electrostatic and first-order exchange interaction energies as defined in symmetry-adapted perturbation theory (SAPT) but computed using monomer Kohn-Sham (KS) determinants, and (ii) the induction energy computed using the coupled KS static response theory, (iii) the exchange-induction and exchange-dispersion energies computed using KS orbitals and orbital energies, the resulting method, denoted by SAPT(DFT), produces very accurate total interaction potentials. For the helium dimer, the only system with nearly exact benchmark values, SAPT(DFT) reproduces the interaction energy to within about 2% at the min. and to a similar accuracy for all other distances ranging from the strongly repulsive to the asymptotic region. For the remaining systems investigated by us, the quality of the SAPT(DFT) interaction energies is so high that these energies may actually be more accurate than the best available results obtained with wave-function techniques. At the same time, SAPT(DFT) is much more computationally efficient than any method previously used for calcg. the dispersion and other interaction energy components at this level of accuracy.
- 316Hesselmann, A.; Jansen, G.; Schütz, M. Interaction Energy Contributions of H-Bonded and Stacked Structures of the AT and GC DNA Base Pairs from the Combined Density Functional Theory and Intermolecular Perturbation Theory Approach J. Am. Chem. Soc. 2006, 128 (36) 11730– 11731 DOI: 10.1021/ja0633363There is no corresponding record for this reference.
- 317Rezáč, J.; Riley, K. E.; Hobza, P. S66: A Well-Balanced Database of Benchmark Interaction Energies Relevant to Biomolecular Structures J. Chem. Theory Comput. 2011, 7 (8) 2427– 2438 DOI: 10.1021/ct2002946317S66: A Well-balanced Database of Benchmark Interaction Energies Relevant to Biomolecular StructuresRezac, Jan; Riley, Kevin E.; Hobza, PavelJournal of Chemical Theory and Computation (2011), 7 (8), 2427-2438CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)With numerous new quantum chem. methods being developed in recent years and the promise of even more new methods to be developed in the near future, it is clearly crit. that highly accurate, well-balanced, ref. data for many different at. and mol. properties be available for the parametrization and validation of these methods. One area of research that is of particular importance in many areas of chem., biol., and material science is the study of noncovalent interactions. Because these interactions are often strongly influenced by correlation effects, it is necessary to use computationally expensive high-order wave function methods to describe them accurately. Here, the authors present a large new database of interaction energies calcd. using an accurate CCSD(T)/CBS scheme. Data are presented for 66 mol. complexes, at their ref. equil. geometries and at 8 points systematically exploring their dissocn. curves; in total, the database contains 594 points: 66 at equil. geometries, and 528 in dissocn. curves. The data set is designed to cover the most common types of noncovalent interactions in biomols., while keeping a balanced representation of dispersion and electrostatic contributions. The data set is therefore well suited for testing and development of methods applicable to bioorg. systems. In addn. to the benchmark CCSD(T) results, the authors also provide decompns. of the interaction energies by DFT-SAPT calcns. The data set was used to test several correlated QM methods, including those parametrized specifically for noncovalent interactions. Among these, the SCS-MI-CCSD method outperforms all other tested methods, with a root-mean-square error of 0.08 kcal/mol for the S66 data set.
- 318Sinnokrot, M. O.; Sherrill, C. D. Substituent Effects in Π–π Interactions: Sandwich and T-Shaped Configurations J. Am. Chem. Soc. 2004, 126 (24) 7690– 7697 DOI: 10.1021/ja049434a318Substituent Effects in π-π Interactions: Sandwich and T-Shaped ConfigurationsSinnokrot, Mutasem Omar; Sherrill, C. DavidJournal of the American Chemical Society (2004), 126 (24), 7690-7697CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Sandwich and T-shaped configurations of benzene dimer, benzene-phenol, benzene-toluene, benzene-fluorobenzene, and benzene-benzonitrile are studied by coupled-cluster theory to elucidate how substituents tune π-π interactions. All substituted sandwich dimers bind more strongly than benzene dimer, whereas the T-shaped configurations bind more or less favorably depending on the substituent. Symmetry-adapted perturbation theory (SAPT) indicates that electrostatic, dispersion, induction, and exchange-repulsion contributions are all significant to the overall binding energies, and all but induction are important in detg. relative energies. Models of π-π interactions based solely on electrostatics, such as the Hunter-Sanders rules, do not seem capable of explaining the energetic ordering of the dimers considered.
- 319Sure, R.; Grimme, S. Corrected Small Basis Set Hartree-Fock Method for Large Systems J. Comput. Chem. 2013, 34 (19) 1672– 1685 DOI: 10.1002/jcc.23317319Corrected small basis set Hartree-Fock method for large systemsSure, Rebecca; Grimme, StefanJournal of Computational Chemistry (2013), 34 (19), 1672-1685CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A quantum chem. method based on a Hartree-Fock calcn. with a small Gaussian AO basis set is presented. Its main area of application is the computation of structures, vibrational frequencies, and noncovalent interaction energies in huge mol. systems. The method is suggested as a partial replacement of semiempirical approaches or d. functional theory (DFT) in particular when self-interaction errors are acute. In order to get accurate results three phys. plausible atom pair-wise correction terms are applied for London dispersion interactions (D3 scheme), basis set superposition error (gCP scheme), and short-ranged basis set incompleteness effects. In total nine global empirical parameters are used. This so-called Hartee-Fock-3c (HF-3c) method is tested for geometries of small org. mols., interaction energies and geometries of noncovalently bound complexes, for supramol. systems, and protein structures. In the majority of realistic test cases good results approaching large basis set DFT quality are obtained at a tiny fraction of computational cost. © 2013 Wiley Periodicals, Inc.
- 320Goerigk, L.; Collyer, C. A.; Reimers, J. R. Recommending Hartree–Fock Theory with London-Dispersion and Basis-Set-Superposition Corrections for the Optimization or Quantum Refinement of Protein Structures J. Phys. Chem. B 2014, 118 (50) 14612– 14626 DOI: 10.1021/jp510148hThere is no corresponding record for this reference.
- 321Brandenburg, J.; Grimme, S. Dispersion Corrected Hartree-Fock and Density Functional Theory for Organic Crystal Structure Prediction Top. Curr. Chem. 2013, 345, 1 DOI: 10.1007/128_2013_488There is no corresponding record for this reference.
- 322Brandenburg, J. G.; Hochheim, M.; Bredow, T.; Grimme, S. Low-Cost Quantum Chemical Methods for Noncovalent Interactions J. Phys. Chem. Lett. 2014, 5 (24) 4275– 4284 DOI: 10.1021/jz5021313322Low-Cost Quantum Chemical Methods for Noncovalent InteractionsBrandenburg, Jan Gerit; Hochheim, Manuel; Bredow, Thomas; Grimme, StefanJournal of Physical Chemistry Letters (2014), 5 (24), 4275-4284CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)A review. The efficient and reasonably accurate description of noncovalent interactions is important for various areas of chem., ranging from supramol. host-guest complexes and biomol. applications to the challenging task of crystal structure prediction. While London dispersion inclusive d. functional theory (DFT-D) can be applied, faster "low-cost" methods are required for large-scale applications. In this Perspective, we present the state-of-the-art of minimal basis set, semiempirical mol.-orbital-based methods. Various levels of approxns. are discussed based either on canonical Hartree-Fock or on semilocal d. functionals. The performance for intermol. interactions is examd. on various small to large mol. complexes and org. solids covering many different chem. groups and interaction types. We put the accuracy of low-cost methods into perspective by comparing with first-principle d. functional theory results. The mean unsigned deviations of binding energies from ref. data are typically 10-30%, which is only two times larger than those of DFT-D. In particular, for neutral or moderately polar systems, many of the tested methods perform very well, while at the same time, computational savings of up to 2 orders of magnitude can be achieved.
- 323Grimme, S.; Brandenburg, J. G.; Bannwarth, C.; Hansen, A. Consistent Structures and Interactions by Density Functional Theory with Small Atomic Orbital Basis Sets J. Chem. Phys. 2015, 143 (5) 054107 DOI: 10.1063/1.4927476323Consistent structures and interactions by density functional theory with small atomic orbital basis setsGrimme, Stefan; Brandenburg, Jan Gerit; Bannwarth, Christoph; Hansen, AndreasJournal of Chemical Physics (2015), 143 (5), 054107/1-054107/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A d. functional theory (DFT) based composite electronic structure approach is proposed to efficiently compute structures and interaction energies in large chem. systems. It is based on the well-known and numerically robust Perdew-Burke-Ernzerhoff (PBE) generalized-gradient-approxn. in a modified global hybrid functional with a relatively large amt. of non-local Fock-exchange. The orbitals are expanded in Ahlrichs-type valence-double zeta AO Gaussian basis sets, which are available for many elements. In order to correct for the basis set superposition error (BSSE) and to account for the important long-range London dispersion effects, our well-established atom-pairwise potentials are used. In the design of the new method, particular attention has been paid to an accurate description of structural parameters in various covalent and non-covalent bonding situations as well as in periodic systems. Together with the recently proposed three-fold cor. (3c) Hartree-Fock method, the new composite scheme (termed PBEh-3c) represents the next member in a hierarchy of "low-cost" electronic structure approaches. They are mainly free of BSSE and account for most interactions in a phys. sound and asymptotically correct manner. PBEh-3c yields good results for thermochem. properties in the huge GMTKN30 energy database. Furthermore, the method shows excellent performance for non-covalent interaction energies in small and large complexes. For evaluating its performance on equil. structures, a new compilation of std. test sets is suggested. These consist of small (light) mols., partially flexible, medium-sized org. mols., mols. comprising heavy main group elements, larger systems with long bonds, 3d-transition metal systems, non-covalently bound complexes (S22 and S66×8 sets), and peptide conformations. For these sets, overall deviations from accurate ref. data are smaller than for various other tested DFT methods and reach that of triple-zeta AO basis set second-order perturbation theory (MP2/TZ) level at a tiny fraction of computational effort. Periodic calcns. conducted for mol. crystals to test structures (including cell vols.) and sublimation enthalpies indicate very good accuracy competitive to computationally more involved plane-wave based calcns. PBEh-3c can be applied routinely to several hundreds of atoms on a single processor and it is suggested as a robust "high-speed" computational tool in theor. chem. and physics. (c) 2015 American Institute of Physics.
- 324Stewart, J. J. Optimization of Parameters for Semiempirical Methods I. Method J. Comput. Chem. 1989, 10 (2) 209– 220 DOI: 10.1002/jcc.540100208324Optimization of parameters for semiempirical methods. I. MethodStewart, James J. P.Journal of Computational Chemistry (1989), 10 (2), 209-20CODEN: JCCHDD; ISSN:0192-8651.A new method for obtaining optimized parameters for semiempirical methods is developed and applied to the MNDO method. The method uses derivs. of calcd. values for properties with respect to adjustable parameters to obtain the optimized values of parameters. The large increase in speed is a result of using a simple series expression for calcd. values of properties rather than employing full semiempirical calcns. With this optimization procedure, the rate-detg. step for parameterizing elements changes from the mechanics of parameterization to the assembling of exptl. ref. data.
- 325Weber, W.; Thiel, W. Orthogonalization Corrections for Semiempirical Methods Theor. Chem. Acc. 2000, 103 (6) 495– 506 DOI: 10.1007/s002149900083325Orthogonalization corrections for semiempirical methodsWeber, Wolfgang; Thiel, WalterTheoretical Chemistry Accounts (2000), 103 (6), 495-506CODEN: TCACFW; ISSN:1432-881X. (Springer-Verlag)Based on a general discussion of orthogonalization effects, two new one-electron orthogonalization corrections are derived to improve existing semiempirical models at the NDDO level. The first one accounts for valence-shell orthogonalization effects on the resonance integrals, while the second one includes the dominant repulsive core-valence interactions through an effective core potential. The corrections for the resonance integrals consist of three-center terms which incorporate stereodiscriminating properties into the two-center matrix elements of the core Hamiltonian. They provide a better description of conformational properties, which is rationalized qual. and demonstrated through numerical calcns. on small model systems. The proposed corrections are part of a new general-purpose semiempirical method which will be described elsewhere.
- 326Stewart, J. J. Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters J. Mol. Model. 2013, 19 (1) 1– 32 DOI: 10.1007/s00894-012-1667-x326Optimization of parameters for semiempirical methods VI: more modifications to the NDDO approximations and re-optimization of parametersStewart, James J. P.Journal of Molecular Modeling (2013), 19 (1), 1-32CODEN: JMMOFK; ISSN:0948-5023. (Springer)Modern semiempirical methods are of sufficient accuracy when used in the modeling of mols. of the same type as used as ref. data in the parameterization. Outside that subset, however, there is an abundance of evidence that these methods are of very limited utility. In an attempt to expand the range of applicability, a new method called PM7 has been developed. PM7 was parameterized using exptl. and high-level ab initio ref. data, augmented by a new type of ref. data intended to better define the structure of parameter space. The resulting method was tested by modeling crystal structures and heats of formation of solids. Two changes were made to the set of approxns.: a modification was made to improve the description of noncovalent interactions, and two minor errors in the NDDO formalism were rectified. Av. unsigned errors (AUEs) in geometry and ΔH f for PM7 were reduced relative to PM6; for simple gas-phase org. systems, the AUE in bond lengths decreased by about 5 % and the AUE in ΔH f decreased by about 10 %; for org. solids, the AUE in ΔH f dropped by 60 % and the redn. was 33.3 % for geometries. A two-step process (PM7-TS) for calcg. the heights of activation barriers has been developed. Using PM7-TS, the AUE in the barrier heights for simple org. reactions was decreased from values of 12.6 kcal/mol-1 in PM6 and 10.8 kcal/mol-1 in PM7 to 3.8 kcal/mol-1. The origins of the errors in NDDO methods have been examd., and were found to be attributable to inadequate and inaccurate ref. data. This conclusion provides insight into how these methods can be improved.
- 327Tuttle, T.; Thiel, W. OM X-D: Semiempirical Methods with Orthogonalization and Dispersion Corrections. Implementation and Biochemical Application Phys. Chem. Chem. Phys. 2008, 10 (16) 2159– 2166 DOI: 10.1039/b718795eThere is no corresponding record for this reference.
- 328Thiel, W. Semiempirical Quantum–chemical Methods Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4 (2) 145– 157 DOI: 10.1002/wcms.1161328Semiempirical quantum-chemical methodsThiel, WalterWiley Interdisciplinary Reviews: Computational Molecular Science (2014), 4 (2), 145-157CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)The semiempirical methods of quantum chem. are reviewed, with emphasis on established NDDO-based methods (MNDO, AM1, PM3) and on the more recent orthogonalization-cor. methods (OM1, OM2, OM3). After a brief historical overview, the methodol. is presented in nontech. terms, covering the underlying concepts, parameterization strategies, and computational aspects, as well as linear scaling and hybrid approaches. The application section addresses selected recent benchmarks and surveys ground-state and excited-state studies, including recent OM2-based excited-state dynamics investigations.
- 329Řezáč, J.; Fanfrlík, J.; Salahub, D.; Hobza, P. Semiempirical Quantum Chemical PM6 Method Augmented by Dispersion and H-Bonding Correction Terms Reliably Describes Various Types of Noncovalent Complexes J. Chem. Theory Comput. 2009, 5 (7) 1749– 1760 DOI: 10.1021/ct9000922329Semiempirical Quantum Chemical PM6 Method Augmented by Dispersion and H-Bonding Correction Terms Reliably Describes Various Types of Noncovalent ComplexesRezac, Jan; Fanfrlik, Jindrich; Salahub, Dennis; Hobza, PavelJournal of Chemical Theory and Computation (2009), 5 (7), 1749-1760CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Because of its construction and parametrization for more than 80 elements, the semiempirical quantum chem. PM6 method is superior to other similar methods. Despite its advantages, however, the PM6 method fails for the description of noncovalent interactions, specifically the dispersion energy and H-bonding. Upon inclusion of correction terms for dispersion and H-bonding, the performance of the method was found to be dramatically improved. The former correction included two parameters in the damping function that were parametrized to reproduce the benchmark interaction energies [CCSD(T)/complete basis set (CBS) limit] of the dispersion-bonded complexes from the S22 data set. The latter correction was parametrized on an extended set of H-bonded stabilization energies detd. at the MP2/cc-pVTZ level. The resulting PM6-DH method was tested on the S22 data set, for which chem. accuracy (error < 1 kcal/mol) was achieved, and also on the JSCH2005 set, for which significant improvement over the original PM6 method was also obtained. Implementation of anal. gradients allows very efficient geometry optimization, which, for all complexes, provides better agreement with the benchmark data. Excellent results were also achieved for small peptides, and here again, chem. accuracy was obtained (i.e., the error with respect to CCSD(T)/CBS results was smaller than 1 kcal/mol). The performance of the technique was finally demonstrated on extended complexes, namely, the porphine dimer and various graphene models with DNA bases and base pairs, where the PM6-DH stabilization energies agree very well with available benchmark data obtained with DFT-D, SCS-MP2, and MP2.5 methods. The PM6-DH calcns. are very efficient and can be routinely applied for systems of up to 1000 atoms. For nonarom. systems, the use of a linear scaling version of the SCF procedure based on localized orbitals speeds up the method significantly and allows one to investigate systems with several thousand atoms. The method can thus replace force fields, which face basic problems for the description of quantum effects, in many applications.
- 330Korth, M.; Pitoňák, M.; Řezáč, J.; Hobza, P. A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical Methods J. Chem. Theory Comput. 2010, 6 (1) 344– 352 DOI: 10.1021/ct900541n330A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical MethodsKorth, Martin; Pitonak, Michal; Rezac, Jan; Hobza, PavelJournal of Chemical Theory and Computation (2010), 6 (1), 344-352CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Semiempirical methods could offer a feasible compromise between ab initio and empirical approaches for the calcn. of large mols. with biol. relevance. A key problem for attempts in this direction is the rather bad performance of current semiempirical methods for noncovalent interactions, esp. hydrogen-bonding. On the basis of the recently introduced PM6-DH method, which includes empirical corrections for dispersion (D) and hydrogen-bond (H) interactions, we have developed an improved and transferable H-bonding correction for semiempirical quantum chem. methods. The performance of the improved correction is evaluated for PM6, AM1, OM3, and SCC-DFTB (enhanced by std. empirical dispersion corrections) with several test sets for noncovalent interactions and is shown to reach the quality of current DFT-D approaches for these types of problems.
- 331Porezag, D.; Frauenheim, T.; Köhler, T.; Seifert, G.; Kaschner, R. Construction of Tight-Binding-like Potentials on the Basis of Density-Functional Theory: Application to Carbon Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 51 (19) 12947 DOI: 10.1103/PhysRevB.51.12947331Construction of tight-binding-like potentials on the basis of density-functional theory: application to carbonPorezag, D.; Frauenheim, Th.; Koehler, Th.; Seifert, G.; Kaschner, R.Physical Review B: Condensed Matter (1995), 51 (19), 12947-57CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)The authors present a d.-functional-based scheme for detg. the necessary parameters of common nonorthogonal tight-binding (TB) models within the framework of the LCAO formalism using the local-d. approxn. (LDA). By only considering two-center integrals the Hamiltonian and overlap matrix elements are calcd. out of suitable input densities and potentials rather than fitted to exptl. data. Anal. functions can be derived for the C-C, C-H, and H-H Hamiltonians and overlap matrix elements. The usual short-range repulsive potential appearing in most TB models is fitted to self-consistent calcns. performed within the LDA. The calcn. of forces is easy and allows an application of the method to mol.-dynamics simulations. Despite its extreme simplicity, the method is transferable to complex carbon and hydrocarbon systems. The detn. of equil. geometries, total energies, and vibrational modes of carbon clusters, hydrocarbon mols., and solid-state modifications of carbon yield results showing an overall good agreement with more sophisticated methods.
- 332Seifert, G.; Porezag, D.; Frauenheim, T. Calculations of Molecules, Clusters, and Solids with a Simplified LCAO-DFT-LDA Scheme Int. J. Quantum Chem. 1996, 58 (2) 185– 192 DOI: 10.1002/(SICI)1097-461X(1996)58:2<185::AID-QUA7>3.3.CO;2-BThere is no corresponding record for this reference.
- 333Elstner, M.; Hobza, P.; Frauenheim, T.; Suhai, S.; Kaxiras, E. Hydrogen Bonding and Stacking Interactions of Nucleic Acid Base Pairs: A Density-Functional-Theory Based Treatment J. Chem. Phys. 2001, 114 (12) 5149– 5155 DOI: 10.1063/1.1329889333Hydrogen bonding and stacking interactions of nucleic acid base pairs: A density-functional-theory based treatmentElstner, Marcus; Hobza, Pavel; Frauenheim, Thomas; Suhai, Sandor; Kaxiras, EfthimiosJournal of Chemical Physics (2001), 114 (12), 5149-5155CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We extend an approx. d. functional theory (DFT) method for the description of long-range dispersive interactions which are normally neglected by construction, irresp. of the correlation function applied. An empirical formula, consisting of an R-6 term is introduced, which is appropriately damped for short distances; the corresponding C6 coeff., which is calcd. from exptl. at. polarizabilities, can be consistently added to the total energy expression of the method. We apply this approx. DFT plus dispersion energy method to describe the hydrogen bonding and stacking interactions of nucleic acid base pairs. Comparison to MP2/6-31G*(0.25) results shows that the method is capable of reproducing hydrogen bonding as well as the vertical and twist dependence of the interaction energy very accurately.
- 334Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, T.; Suhai, S.; Seifert, G. Self-Consistent-Charge Density-Functional Tight-Binding Method for Simulations of Complex Materials Properties Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58 (11) 7260 DOI: 10.1103/PhysRevB.58.7260334Self-consistent-charge density-functional tight-binding method for simulations of complex materials propertiesElstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, Th.; Suhai, S.; Seifert, G.Physical Review B: Condensed Matter and Materials Physics (1998), 58 (11), 7260-7268CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)We outline details about an extension of the tight-binding (TB) approach to improve total energies, forces, and transferability. The method is based on a second-order expansion of the Kohn-Sham total energy in d.-functional theory (DFT) with respect to charge-d. fluctuations. The zeroth-order approach is equiv. to a common std. non-self-consistent (TB) scheme, while at second-order a transparent, parameter-free, and readily calculable expression for generalized Hamiltonian matrix elements may be derived. These are modified by a self-consistent redistribution of Mulliken charges (SCC). Besides the usual "band structure" and short-range repulsive terms the final approx. Kohn-Sham energy addnl. includes a Coulomb interaction between charge fluctuations. At large distances this accounts for long-range electrostatic forces between two point charges and approx. includes self-interaction contributions of a given atom if the charges are located at one and the same atom. We apply the new SCC scheme to problems where deficiencies within the non-SCC std. TB approach become obvious. We thus considerably improve transferability.
- 335Kunsági-Máté, S.; Ortmann, E.; Kollár, L.; Szabó, K.; Nikfardjam, M. P. Effect of Ferrous and Ferric Ions on Copigmentation in Model Solutions J. Mol. Struct. 2008, 891 (1–3) 471– 474 DOI: 10.1016/j.molstruc.2008.04.036There is no corresponding record for this reference.
- 336Bondi, A. J. Van Der Waals Volumes and Radii J. Phys. Chem. 1964, 68 (3) 441– 451 DOI: 10.1021/j100785a001336van der Waals volumes and radiiBondi, A.Journal of Physical Chemistry (1964), 68 (3), 441-51CODEN: JPCHAX; ISSN:0022-3654.Intermol. van der Waals radii of the nonmetallic elements were assembled into a list of recommended values for vol. calcns. These values were arrived at by selecting from the most reliable x-ray diffraction data those which could be reconciled with crystal d. at 0°K. (to give reasonable packing d.), gas kinetic collision cross section, crit. d., and with liquid state properties. A qual. understanding of the nature of van der Waals radii is provided by correlation with the de Broglie wavelength of the outermost valence electron. Tentative values for the van der Waals radii of metallic elements in organometallic compds. are proposed. A list of increments for the vol. of mols. impenetrable to thermal collision, the so-called van der Waals vol., and of the corresponding increments in area per mol. is given.
- 337
As seen in previous sections, we are aware that in principle, CCSD(T), or other methods such as QMC, are more desirable. However, such methods are currently too computationally demanding to be applied to real-world copigmentation systems.
There is no corresponding record for this reference. - 338Freitas, A. A.; Shimizu, K.; Dias, L. G.; Quina, F. H. A Computational Study of Substituted Flavylium Salts and Their Quinonoidal Conjugate-Bases: S0→S1 Electronic Transition, Absolute pKa and Reduction Potential Calculations by DFT and Semiempirical Methods J. Braz. Chem. Soc. 2007, 18 (8) 1537– 1546 DOI: 10.1590/S0103-50532007000800014338A computational study of substituted flavylium salts and their quinonoidal conjugate-bases: S0→S1 electronic transition, absolute pKa and reduction potential calculations by DFT and semiempirical methodsFreitas, Adilson A.; Shimizu, K.; Dias, Luis G.; Quina, Frank H.Journal of the Brazilian Chemical Society (2007), 18 (8), 1537-1546CODEN: JOCSET; ISSN:0103-5053. (Sociedade Brasileira de Quimica)The electronic transitions for flavylium cations and quinonoidal bases of 17 substituted flavylium salts have been studied at semiempirical and DFT (d. functional theory) levels. Solvent effect on electronic spectra was included by Polarizable Continuum Model, PCM. We assigned longest-wavelength absorption maxima to HOMO→LUMO transition. Both levels of theory gave good results for electronic transitions of flavylium cations whereas only TDDFT-PCM calcns. could be used for electronic transitions of their quinonoidal bases. We also performed abs. pKa calcns. of nine flavylium salts at DFT level. The pKa calcd. values by our PCM parameterization gave excellent results with mean abs. deviation less than a half of one pKa unit. One-electron redn. potentials were carried out for 5 flavylium cations at DFT level. The theor. results found were in good agreement with exptl. values after adjustment for a systematic deviation.
- 339Sakata, K.; Saito, N.; Honda, T. Ab Initio Study of Molecular Structures and Excited States in Anthocyanidins Tetrahedron 2006, 62 (15) 3721– 3731 DOI: 10.1016/j.tet.2006.01.081339Ab initio study of molecular structures and excited states in anthocyanidinsSakata, Ken; Saito, Norio; Honda, ToshioTetrahedron (2006), 62 (15), 3721-3731CODEN: TETRAB; ISSN:0040-4020. (Elsevier B.V.)The structural and electronic characters of four types of hydroxyl group-substituted anthocyanidins (pelargonidin, cyanidin, delphinidin, and aurantinidin) were examd. using quantum chem. calcns. For these cationic mols., both the planar and non-planar structures in the electronic ground state were detd. at the B3LYP/D95 level of theory. We revealed that the planar structure is slightly more stable than the non-planar structure for each mol. For the optimized planar structures, single excitation-CI (SE-CI) based on the RHF (RHF) wave function was evaluated and the electronic character in the low-excited states was discussed in terms of the MO theory. Symmetry adapted cluster (SAC)/SAC-CI calcns. were also carried out to est. the excitation energies precisely. The results showed that hydroxylation of the Ph group causes a change in the excitation energies without taking the solvent effects into account. The results are in agreement with spectral expts. and previous MO calcns.
- 340Anouar, E. H.; Gierschner, J.; Duroux, J.-L.; Trouillas, P. UV/Visible Spectra of Natural Polyphenols: A Time-Dependent Density Functional Theory Study Food Chem. 2012, 131 (1) 79– 89 DOI: 10.1016/j.foodchem.2011.08.034340UV/Visible spectra of natural polyphenols: A time-dependent density functional theory studyAnouar, El Hassane; Gierschner, Johannes; Duroux, Jean-Luc; Trouillas, PatrickFood Chemistry (2012), 131 (1), 79-89CODEN: FOCHDJ; ISSN:0308-8146. (Elsevier Ltd.)In addn. to their numerous biol. activities, natural and hemisynthetic polyphenols contribute to the large variety of colors (from red to violet) in nature (e.g., fruit, vegetables, leaves and petals). In order to understand the color variation attributed to the multitude of chem. structures of this wide class of compds., time-dependent d. functional quantum-chem. calcns. at the B3P86/6-311+G(d,p) level of theory appears as a relevant and efficient tool. The UV/Vis properties of 33 polyphenols were systematically investigated, including mainly flavonoids, isoflavonoids and flavonolignans. On the basis of MO anal. we established the structure-property relationship, inter alia showing the role of π orbital (de-)localization, mesomeric (+M) effects of hydroxyl groups and structural modification of the mol. backbone. The results might help in the future, for example, for the prediction of novel hemisynthetic compds.
- 341Millot, M.; Di Meo, F.; Tomasi, S.; Boustie, J.; Trouillas, P. Photoprotective Capacities of Lichen Metabolites: A Joint Theoretical and Experimental Study. J. Photochem. Photobiol., B 111, 17– 26. DOI: 10.1016/j.jphotobiol.2012.03.005There is no corresponding record for this reference.
- 342Carvalho, A. R. F.; Oliveira, J.; De Freitas, V.; Mateus, N.; Melo, A. Unusual Color Change of Vinylpyranoanthocyanin–Phenolic Pigments J. Agric. Food Chem. 2010, 58 (7) 4292– 4297 DOI: 10.1021/jf904246gThere is no corresponding record for this reference.
- 343Quartarolo, A. D.; Russo, N. A Computational Study (TDDFT and RICC2) of the Electronic Spectra of Pyranoanthocyanins in the Gas Phase and Solution J. Chem. Theory Comput. 2011, 7 (4) 1073– 1081 DOI: 10.1021/ct2000974343A Computational Study (TDDFT and RICC2) of the Electronic Spectra of Pyranoanthocyanins in the Gas Phase and SolutionQuartarolo, Angelo Domenico; Russo, NinoJournal of Chemical Theory and Computation (2011), 7 (4), 1073-1081CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The conformational structures and UV-vis absorption electronic spectra of a class of derived anthocyanin mols. (pyranoanthocyanins) have been investigated mainly by means of d. functional (DFT) and time-dependent DFT methods. Pyranoanthocyanins are natural pigments present in aged wines and absorb at shorter wavelengths (around 500 nm) than the parent anthocyanin compds., giving an orange-brown colored soln. The investigated mols. are derived from the reaction of glycosylated malvidin, peonidin, and petunidin with enolizable mols. (acetaldehyde and pyruvic acid) and vinyl derivs. During wine storage, the concn. of pyranoanthocyanins increases with time, and anal. measurements (e.g., UV-vis spectroscopy) can characterize aged wines by color anal. The prediction of absorption electronic spectra from TDDFT results, with the inclusion of water bulk solvation effects through the conductor-like polarizable continuum model, gives an abs. mean deviation from exptl. absorption maxima of 0.1 eV and a good reprodn. of the spectra line shape over the visible range of the spectrum. TDDFT calcd. excitation energies agree with those obtained from ab initio multireference coupled cluster with the resoln. of identity approxn. (RICC2) methods, calcd. at DFT gas-phase geometries.
- 344Milián-Medina, B.; Gierschner, J. Computational Design of Low Singlet–triplet Gap All-Organic Molecules for OLED Application Org. Electron. 2012, 13 (6) 985– 991 DOI: 10.1016/j.orgel.2012.02.010344Computational design of low singlet-triplet gap all-organic molecules for OLED applicationMilian-Medina, Begona; Gierschner, JohannesOrganic Electronics (2012), 13 (6), 985-991CODEN: OERLAU; ISSN:1566-1199. (Elsevier B.V.)It was recently reported that external quantum efficiency in org. LEDs can be substantially enhanced when triplet excitons are harvested through upconversion by E-type delayed fluorescence in materials with small singlet-triplet energy gap ΔEST, based on donor-acceptor (DA) chromophores. Furthermore, org. solar cells (OSCs) might profit from such materials in order to reduce recombination losses. However, targeted design rules for such materials are missing up to now. In this paper, we follow a facile (TD-)DFT-based computational design concept by engineering the fragment frontier orbitals in DA systems. The calcns. show that optimized systems with very small ΔEST in the range of kT can be achieved by balancing the energetic offset between fragment MOs as well as through the nature of the DA connector. Application in OLED will addnl. require small non-radiative rates, which recommends large bandgap materials. Utilization in polymeric DA systems with small ΔEST in OSCs requires the full exploration of the chain length dependence of the resp. oligomers.
- 345Dreuw, A.; Weisman, J. L.; Head-Gordon, M. Long-Range Charge-Transfer Excited States in Time-Dependent Density Functional Theory Require Non-Local Exchange J. Chem. Phys. 2003, 119 (6) 2943– 2946 DOI: 10.1063/1.1590951345Long-range charge-transfer excited states in time-dependent density functional theory require non-local exchangeDreuw, Andreas; Weisman, Jennifer L.; Head-Gordon, MartinJournal of Chemical Physics (2003), 119 (6), 2943-2946CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The electrostatic attraction between the sepd. charges in long-range excited charge-transfer states originates from the non-local Hartree-Fock exchange potential and is, thus, a non-local property. Present-day time-dependent d. functional theory employing local exchange-correlation functionals does not capture this effect and therefore fails to describe charge-transfer excited states correctly. A hybrid method that is qual. correct is described.
- 346Dreuw, A.; Head-Gordon, M. Single-Reference Ab Initio Methods for the Calculation of Excited States of Large Molecules Chem. Rev. 2005, 105 (11) 4009– 4037 DOI: 10.1021/cr0505627346Single-Reference ab Initio Methods for the Calculation of Excited States of Large MoleculesDreuw, Andreas; Head-Gordon, MartinChemical Reviews (Washington, DC, United States) (2005), 105 (11), 4009-4037CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review of single-ref. ab initio excited state methods, which are applicable to large mols. and do not explicitly include correlation through the ground-state wave function. CIS, TDHF, and TDDFT are rigorously introduced by outlining their derivations and theor. footing, where special emphasis is put on their relations to each other. Different methods for the anal. of complicated electronically excited states are reviewed and comparatively discussed. The applicability of the presented methods is outlined, their limitations are show, and out their strengths and weaknesses are pointed out.
- 347Casida, M. E.; Gutierrez, F.; Guan, J.; Gadea, F.-X.; Salahub, D.; Daudey, J.-P. Charge-Transfer Correction for Improved Time-Dependent Local Density Approximation Excited-State Potential Energy Curves: Analysis within the Two-Level Model with Illustration for H2 and LiH J. Chem. Phys. 2000, 113 (17) 7062– 7071 DOI: 10.1063/1.1313558347Charge-transfer correction for improved time-dependent local density approximation excited-state potential energy curves: Analysis within the two-level model with illustration for H2 and LiHCasida, Mark E.; Gutierrez, Fabien; Guan, Jingang; Gadea, Florent-Xavier; Salahub, Dennis; Daudey, Jean-PierreJournal of Chemical Physics (2000), 113 (17), 7062-7071CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Time-dependent d.-functional theory (TDDFT) is an increasingly popular approach for calcg. mol. excitation energies. However, the TDDFT lowest triplet excitation energy, ωT, of a closed-shell mol. often falls rapidly to zero and then becomes imaginary at large internuclear distances. We show that this unphys. behavior occurs because ωT2 must become neg. wherever symmetry breaking lowers the energy of the ground state soln. below that of the symmetry unbroken soln. We use the fact that the ΔSCF method gives a qual. correct first triplet excited state to derive a "charge-transfer correction" (CTC) for the time-dependent local d. approxn. (TDLDA) within the two-level model and the Tamm-Dancoff approxn. (TDA). Although this correction would not be needed for the exact exchange-correlation functional, it is evidently important for a correct description of mol. excited state potential energy surfaces in the TDLDA. As a byproduct of our anal., we show why TDLDA and LDA ΔSCF excitation energies are often very similar near the equil. geometries. The reasoning given here is fairly general and it is expected that similar corrections will be needed in the case of generalized gradient approxns. and hybrid functionals.
- 348Yanai, T.; Tew, D. P.; Handy, N. C. A New Hybrid Exchange–correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP) Chem. Phys. Lett. 2004, 393 (1–3) 51– 57 DOI: 10.1016/j.cplett.2004.06.011348A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP)Yanai, Takeshi; Tew, David P.; Handy, Nicholas C.Chemical Physics Letters (2004), 393 (1-3), 51-57CODEN: CHPLBC; ISSN:0009-2614. (Elsevier Science B.V.)A new hybrid exchange-correlation functional named CAM-B3LYP is proposed. It combines the hybrid qualities of B3LYP and the long-range correction presented by Tawada et al. [J. Chem. Phys., in press]. We demonstrate that CAM-B3LYP yields atomization energies of similar quality to those from B3LYP, while also performing well for charge transfer excitations in a dipeptide model, which B3LYP underestimates enormously. The CAM-B3LYP functional comprises of 0.19 Hartree-Fock (HF) plus 0.81 Becke 1988 (B88) exchange interaction at short-range, and 0.65 HF plus 0.35 B88 at long-range. The intermediate region is smoothly described through the std. error function with parameter 0.33.
- 349Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–atom Dispersion Corrections Phys. Chem. Chem. Phys. 2008, 10 (44) 6615– 6620 DOI: 10.1039/b810189b349Long-range corrected hybrid density functionals with damped atom-atom dispersion correctionsChai, Jeng-Da; Head-Gordon, MartinPhysical Chemistry Chemical Physics (2008), 10 (44), 6615-6620CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)We report re-optimization of a recently proposed long-range cor. (LC) hybrid d. functional [J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106] to include empirical atom-atom dispersion corrections. The resulting functional, ωB97X-D yields satisfactory accuracy for thermochem., kinetics, and non-covalent interactions. Tests show that for non-covalent systems, ωB97X-D shows slight improvement over other empirical dispersion-cor. d. functionals, while for covalent systems and kinetics it performs noticeably better. Relative to our previous functionals, such as ωB97X, the new functional is significantly superior for non-bonded interactions, and very similar in performance for bonded interactions.
- 350Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Double-Hybrid Density Functionals J. Chem. Phys. 2009, 131 (17) 174105 DOI: 10.1063/1.3244209350Long-range corrected double-hybrid density functionalsChai, Jeng-Da; Head-Gordon, MartinJournal of Chemical Physics (2009), 131 (17), 174105/1-174105/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We extend the range of applicability of our previous long-range cor. (LC) hybrid functional, ωB97X, with a nonlocal description of electron correlation, inspired by second-order Moller-Plesset (many-body) perturbation theory. This LC "double-hybrid" d. functional, denoted as ωB97X-2, is fully optimized both at the complete basis set limit (using 2-point extrapolation from calcns. using triple and quadruple zeta basis sets), and also sep. using the somewhat less expensive 6-311++G(3df,3pd) basis. On independent test calcns. (as well as training set results), ωB97X-2 yields high accuracy for thermochem., kinetics, and noncovalent interactions. In addn., owing to its high fraction of exact Hartree-Fock exchange, ωB97X-2 shows significant improvement for the systems where self-interaction errors are severe, such as sym. homonuclear radical cations. (c) 2009 American Institute of Physics.
- 351Caricato, M.; Trucks, G. W.; Frisch, M. J.; Wiberg, K. B. Oscillator Strength: How Does TDDFT Compare to EOM-CCSD? J. Chem. Theory Comput. 2011, 7 (2) 456– 466 DOI: 10.1021/ct100662n351Oscillator Strength: How Does TDDFT Compare to EOM-CCSD?Caricato, Marco; Trucks, Gary W.; Frisch, Michael J.; Wiberg, Kenneth B.Journal of Chemical Theory and Computation (2011), 7 (2), 456-466CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We compare a large variety of d. functionals against the equation of motion coupled cluster singles and doubles (EOM-CCSD) method for the calcn. of oscillator strengths. Valence and Rydberg states are considered for a test set composed of 11 small org. mols. In our previous work, the same systems and methods were tested against exptl. results for the excitation energies. The results from this investigation confirm our previous findings, i.e., that there is a large difference between the functionals. For the oscillator strength, the av. best agreement with EOM-CCSD is provided by CAM-B3LYP followed by LC-ωPBE and, to a lesser extent, B3P86 and LC-BLYP.
- 352Beyhan, S. M.; Götz, A. W.; Ariese, F.; Visscher, L.; Gooijer, C. Computational Study on the Anomalous Fluorescence Behavior of Isoflavones J. Phys. Chem. A 2011, 115 (9) 1493– 1499 DOI: 10.1021/jp109059eThere is no corresponding record for this reference.
- 353Malcıoğlu, O. B.; Calzolari, A.; Gebauer, R.; Varsano, D.; Baroni, S. Dielectric and Thermal Effects on the Optical Properties of Natural Dyes: A Case Study on Solvated Cyanin J. Am. Chem. Soc. 2011, 133 (39) 15425– 15433 DOI: 10.1021/ja201733vThere is no corresponding record for this reference.
- 354Petrone, A.; Cerezo, J.; Ferrer, F. J. A.; Donati, G.; Improta, R.; Rega, N.; Santoro, F. Absorption and Emission Spectral Shapes of a Prototype Dye in Water by Combining Classical/Dynamical and Quantum/Static Approaches J. Phys. Chem. A 2015, 119 (21) 5426– 5438 DOI: 10.1021/jp510838mThere is no corresponding record for this reference.
- 355Sinnecker, S.; Rajendran, A.; Klamt, A.; Diedenhofen, M.; Neese, F. Calculation of Solvent Shifts on Electronic G-Tensors with the Conductor-Like Screening Model (COSMO) and Its Self-Consistent Generalization to Real Solvents (Direct COSMO-RS) J. Phys. Chem. A 2006, 110, 2235– 2245 DOI: 10.1021/jp056016z355Calculation of Solvent Shifts on Electronic g-Tensors with the Conductor-Like Screening Model (COSMO) and Its Self-Consistent Generalization to Real Solvents (Direct COSMO-RS)Sinnecker, Sebastian; Rajendran, Arivazhagan; Klamt, Andreas; Diedenhofen, Michael; Neese, FrankJournal of Physical Chemistry A (2006), 110 (6), 2235-2245CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The conductor-like screening model (COSMO) was used to investigate the solvent influence on electronic g-values of org. radicals. The previously studied di-Ph nitric oxide and di-tert-Bu nitric oxide radicals were taken as test cases. The calcns. employed spin-unrestricted d. functional theory and the BP and B3LYP d. functionals. The g-tensors were calcd. as mixed second deriv. properties with respect to the external magnetic field and the electron magnetic moment. The first-order response of the Kohn-Sham orbitals with respect to the external magnetic field was detd. through the coupled-perturbed DFT approach. The spin-orbit coupling operator was treated using an accurate multicenter spin-orbit mean-field (SOMF) approach. Provided that important hydrogen bonds are explicitly modeled by a supermol. approach and that the basis set is sufficiently satd., the COSMO calcns. lead to accurate predictions of isotropic g-shifts with deviations of not more than 100 ppm relative to expt. Very accurate results were obtained by employing a recently developed self-consistent modification of the COSMO method to real solvents (COSMO-RS), which we briefly introduce in this paper as direct COSMO-RS (D-COSMO-RS). This model gives isotropic g-shifts of similar high accuracy for water without using the supermol. approach. This is an important result because it solves many of the problems assocd. with the supermol. approach such as local min. and the choice of a suitable model system. Thus, the self-consistent D-COSMO-RS incorporates some specific solvation effects into continuum models, in particular it appears to successfully model the effects of hydrogen bonding. Although not yet widely validated, this opens a novel approach for the calcn. of properties which so far only could be calcd. by the inclusion of explicit solvent mols. in continuum solvation methods.
- 356Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models Chem. Rev. 2005, 105, 2999– 3093 DOI: 10.1021/cr9904009356Quantum Mechanical Continuum Solvation ModelsTomasi, Jacopo; Mennucci, Benedetta; Cammi, RobertoChemical Reviews (Washington, DC, United States) (2005), 105 (8), 2999-3093CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review.
- 357Ge, X.; Timrov, I.; Binnie, S.; Biancardi, A.; Calzolari, A.; Baroni, S. Accurate and Inexpensive Prediction of the Color Optical Properties of Anthocyanins in Solution J. Phys. Chem. A 2015, 119 (16) 3816– 3822 DOI: 10.1021/acs.jpca.5b01272There is no corresponding record for this reference.
- 358Pedersen, M. N.; Hedegård, E. D.; Olsen, J. M. H.; Kauczor, J.; Norman, P.; Kongsted, J. Damped Response Theory in Combination with Polarizable Environments: The Polarizable Embedding Complex Polarization Propagator Method J. Chem. Theory Comput. 2014, 10 (3) 1164– 1171 DOI: 10.1021/ct400946k358Damped Response Theory in Combination with Polarizable Environments: The Polarizable Embedding Complex Polarization Propagator MethodPedersen, Morten N.; Hedegaard, Erik D.; Olsen, Jogvan Magnus H.; Kauczor, Joanna; Norman, Patrick; Kongsted, JacobJournal of Chemical Theory and Computation (2014), 10 (3), 1164-1171CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present a combination of the polarizable embedding (PE) scheme with the complex polarization propagator (CPP) method with the aim of calcg. response properties including relaxation for large and complex systems. This new approach, termed PE-CPP, will benefit from the highly advanced description of the environmental electrostatic potential and polarization in the PE method as well as the treatment of near-resonant effects in the CPP approach. The PE-CPP model has been implemented in a Kohn-Sham d. functional theory approach, and we present pilot calcns. exemplifying the implementation for the UV/vis and carbon K-edge X-ray absorption spectra of the protein plastocyanin. Furthermore, tech. details assocd. with a PE-CPP calcn. are discussed.
- 359Olsen, J. M.; Aidas, K.; Kongsted, J. Excited States in Solution through Polarizable Embedding J. Chem. Theory Comput. 2010, 6 (12) 3721– 3734 DOI: 10.1021/ct1003803359Excited States in Solution through Polarizable EmbeddingOlsen, Jogvan Magnus; Aidas, Kestutis; Kongsted, JacobJournal of Chemical Theory and Computation (2010), 6 (12), 3721-3734CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present theory and implementation of an advanced quantum mechanics/mol. mechanics (QM/MM) approach using a fully self-consistent polarizable embedding (PE) scheme. It is a polarizable layered model designed for effective yet accurate inclusion of an anisotropic medium in a quantum mech. calcn. The polarizable embedding potential is described by an atomistic representation including terms up to localized octupoles and anisotropic polarizabilities. It is generally applicable to any quantum chem. description but is here implemented for the case of Kohn-Sham d. functional theory which we denote the PE-DFT method. It has been implemented in combination with time-dependent quantum mech. linear and nonlinear response techniques, thus allowing for assessment of electronic excitation processes and dynamic ground- and excited-state mol. properties using a nonequil. formulation of the environmental response. In our formulation of polarizable embedding we explicitly take into account the full self-consistent many-body environmental response from both ground and excited states. The PE-DFT method can be applied to any mol. system, e.g., proteins, nanoparticles and solute-solvent systems. Here, we present numerical examples of solvent shifts and excited-state properties related to a set of org. mols. in aq. soln.
- 360Sjöqvist, J.; Linares, M.; Mikkelsen, K. V.; Norman, P. QM/MM-MD Simulations of Conjugated Polyelectrolytes: A Study on Luminescent Conjugated Oligothiophenes for Use as Bio-Physical Probes J. Phys. Chem. A 2014, 118 (19) 3419– 3428 DOI: 10.1021/jp5009835There is no corresponding record for this reference.
- 361Avila Ferrer, F. J.; Cerezo, J.; Stendardo, E.; Improta, R.; Santoro, F. Insights for an Accurate Comparison of Computational Data to Experimental Absorption and Emission Spectra: Beyond the Vertical Transition Approximation J. Chem. Theory Comput. 2013, 9 (4) 2072– 2082 DOI: 10.1021/ct301107m361Insights for an Accurate Comparison of Computational Data to Experimental Absorption and Emission Spectra: Beyond the Vertical Transition ApproximationAvila Ferrer, Francisco J.; Cerezo, Javier; Stendardo, Emiliano; Improta, Roberto; Santoro, FabrizioJournal of Chemical Theory and Computation (2013), 9 (4), 2072-2082CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The authors carefully study the relation between computed data and exptl. electronic spectra. To that end, the authors compare both vertical transition energies, EV, and characteristic frequencies of the spectrum like the max., νmax, and the center of gravity, M1, taking advantage of an anal. expression of M1 in terms of the parameters of the initial- and final-state potential energy surfaces. After pointing out that, for an accurate comparison, exptl. spectra should be preliminarily mapped from wavelength to frequency domain and transformed to normalized lineshapes, the authors simulate the absorption and emission spectra of several prototypical chromophores, obtaining lineshapes in very good agreement with exptl. data. Results indicate that the customary comparison of exptl. νmax and computational EV, without taking into account vibrational effects, is not an adequate measure of the performance of an electronic method. In fact, it introduces systematic errors that, in the studied systems, are ∼0.1-0.3 eV, i.e., values comparable to the expected accuracy of the most accurate computational methods. On the contrary, a comparison of exptl. and computed M1 and/or 0-0 transition frequencies provides more robust results. Some rules of thumbs probably help rationalize which kind of correction 1 should expect when comparing EV, M1, and νmax.
- 362Avila Ferrer, F. J.; Cerezo, J.; Soto, J.; Improta, R.; Santoro, F. First-Principle Computation of Absorption and Fluorescence Spectra in Solution Accounting for Vibronic Structure, Temperature Effects and Solvent Inhomogenous Broadening Comput. Theor. Chem. 2014, 1040–1041, 328– 337 DOI: 10.1016/j.comptc.2014.03.003362First-principle computation of absorption and fluorescence spectra in solution accounting for vibronic structure, temperature effects and solvent inhomogeneous broadeningAvila Ferrer, Francisco Jose; Cerezo, Javier; Soto, Juan; Improta, Roberto; Santoro, FabrizioComputational & Theoretical Chemistry (2014), 1040-1041 (), 328-337CODEN: CTCOA5; ISSN:2210-271X. (Elsevier B.V.)We compute the line shape of absorption and emission electronic spectra of two different dyes, coumarin C153 and N-methyl-6-quinolinium betaine accounting for the vibronic structure, temp. effects and polar solvent inhomogeneous broadening, without using any phenomenol. parameter. We exploit a no. of recent developments including a time-dependent (TD) approach to the computation of vibronic spectra that provides fully converged line shapes at finite temp. accounting for both Duschinsky and Herzberg-Teller effects, and the state-specific (SS) implementation of Polarizable Continuum Model (PCM). This latter is adopted to compute the solvent reorganization energy connected to inhomogeneous broadening. We compute the absorption and fluorescence spectra in the gas-phase, non-polar and polar solvents analyzing the relative importance of different sources of broadening. To this end we investigate the performance of time-dependent d. functional theory, complete active space self consistent field (CASSCF), and complete active space 2nd-order perturbation theory (CASPT2) methods in the computation of inhomogeneous broadening.