Lipid14: The Amber Lipid Force Field
- Callum J. Dickson ,
- Benjamin D. Madej ,
- Åge A. Skjevik ,
- Robin M. Betz ,
- Knut Teigen ,
- Ian R. Gould , and
- Ross C. Walker
Abstract

The AMBER lipid force field has been updated to create Lipid14, allowing tensionless simulation of a number of lipid types with the AMBER MD package. The modular nature of this force field allows numerous combinations of head and tail groups to create different lipid types, enabling the easy insertion of new lipid species. The Lennard-Jones and torsion parameters of both the head and tail groups have been revised and updated partial charges calculated. The force field has been validated by simulating bilayers of six different lipid types for a total of 0.5 μs each without applying a surface tension; with favorable comparison to experiment for properties such as area per lipid, volume per lipid, bilayer thickness, NMR order parameters, scattering data, and lipid lateral diffusion. As the derivation of this force field is consistent with the AMBER development philosophy, Lipid14 is compatible with the AMBER protein, nucleic acid, carbohydrate, and small molecule force fields.
Introduction
Parameterization Strategy
Generation of Lipid14 Parameters
Figure 1

Figure 1. A box of 144 pentadecane molecules simulated in the NPT ensemble at 298.15 K using the General Amber Force Field (16) to model the carbon chains.
Hydrocarbon Tail Parameters

LJ parameters | CH2–CH2–CH2–CH2 torsion | |||||
---|---|---|---|---|---|---|
atom type | radius R (Å) | well-depth ε (kcal mol–1) | force constant PK (kcal mol–1) | periodicity PN | phase (deg) | |
Lipid11 | cA | 1.9080 | 0.1094 | 0.20 | 1 | 180 |
hA | 1.4870 | 0.0157 | 0.25 | 2 | 180 | |
0.18 | 3 | 0 | ||||
Lipid14 | cD | 1.9080 | 0.1094 | 0.3112 | 1 | 180 |
hL | 1.4600 | 0.0100 | –0.1233 | 2 | 180 | |
0.1149 | 3 | 0 | ||||
–0.2199 | 4 | 0 | ||||
0.2170 | 5 | 0 |

Figure 2

Figure 2. The energy profile for rotating about selected torsions of a cis-5-decene molecule. Energy evaluated using QM and the HM-IE method (filled triangle ▲), AMBER with standard GAFF parameters (dotted line), and AMBER with Lipid14 parameters (black line). Torsion fits from the top are as follows: CH2–CH–CH–CH2, CH–CH–CH2–CH2, and CH–CH2–CH2–CH2.



Head Group Parameters
LJ parameters | |||||
---|---|---|---|---|---|
atom type | radius R (Å) | well-depth ε (kcal mol–1) | ΔHvap (kJ mol-1) | ρ (kg m-3) | |
Lipid11 | oC | 1.6612 | 0.210 | 39.11 ± 0.04 | 928.38 ± 0.09 |
oS | 1.6837 | 0.170 | |||
cC | 1.9080 | 0.086 | |||
Lipid14 | oC | 1.6500 | 0.140 | 33.0 ± 0.07 | 925.8 ± 0.05 |
oS | 1.6500 | 0.120 | |||
cC | 1.9080 | 0.070 | |||
Expt | 32.29 (41) | 934.2 (41) |
All values at 298.15 K.
Partial Charges
Figure 3

Figure 3. Structure and charges of Lipid11/Lipid14 headgroup and tail group caps. (22)
Head Group Torsion Fits
Figure 4

Figure 4. A capped lauroyl tail group residue was used to fit the oS-cC-cD-cD and oC-cC-cD-cD torsions.
Figure 5

Figure 5. The energy profiles for rotating about selected torsions of a capped lauroyl tail group residue. Energy evaluated using QM and the HM-IE method (filled triangle ▲), AMBER with standard GAFF/Lipid11 parameters (dotted line), and AMBER with Lipid14 parameters (black line). Torsion fits from the top are oC-cC-cD-cD and oS-cC-cD-cD.
Parameterization
Hydrocarbon Parameters
ΔHvap (kJ mol-1) | ρ (kg m-3) | DPBC (10-5 cm2 s-1) | Dcorr (10-5 cm2 s-1) | |
---|---|---|---|---|
Pentane | ||||
Lipid14 | 23.03 ± 0.16 | 592.45 ± 0.16 | 6.45 ± 0.56 | 7.1 ± 0.56 |
Expt | 26.43 (41) | 626.2 (41) | 5.45 (43) | |
Hexane | ||||
Lipid14 | 28.54 ± 0.1 | 636.3 ± 0.09 | 4.55 ± 0.29 | 5.02 ± 0.29 |
Expt | 31.56 (41) | 656, (44) 660.6 (41) | 4.21 (43) | |
Heptane | ||||
Lipid14 | 33.37 ± 0.11 | 667.31 ± 0.14 | 3.47 ± 0.23 | 3.85 ± 0.23 |
Expt | 36.57 (41) | 679.5 (41) | 3.12 (43) | |
Octane | ||||
Lipid14 | 38.67 ± 0.31 | 690.96 ± 0.10 | 2.11 ± 0.15 | 2.46 ± 0.15 |
Expt | 41.49 (41) | 698.6 (41) | 2.354 (45) | |
Decane | ||||
Lipid14 | 49.34 ± 0.30 | 724.47 ± 0.07 | 1.44 ± 0.15 | 1.65 ± 0.15 |
Expt | 51.42 (41) | 726.6 (41) | 1.39 (45) | |
Tridecane | ||||
Lipid14 | 64.62 ± 0.27 | 756.19 ± 0.24 | 0.48 ± 0.04 | 0.57 ± 0.04 |
Expt | 66.68 (41) | 756.4 (41) | 0.712 (45) | |
Pentadecane | ||||
Lipid14 | 74.99 ± 0.39 | 770.67 ± 0.25 | 0.30 ± 0.02 | 0.36 ± 0.02 |
Expt | 76.77 (41) | 768.5 (41) | 0.461 (45) |
All values at 298.15 K.
trans | gauche | t/g ratio | eg | gg | gtg′+gtg | |
---|---|---|---|---|---|---|
Pentane | ||||||
Lipid14 | 1.20 | 0.80 | 1.49 | 0.80 | 0.13 | - |
Hexane | ||||||
Lipid14 | 1.83 | 1.17 | 1.57 | 0.83 | 0.22 | 0.14 |
Heptane | ||||||
Lipid14 | 2.49 | 1.51 | 1.65 | 0.83 | 0.31 | 0.24 |
Octane | ||||||
Lipid14 | 3.15 | 1.85 | 1.71 | 0.81 | 0.39 | 0.33 |
Decane | ||||||
Lipid14 | 4.47 | 2.53 | 1.77 | 0.81 | 0.57 | 0.54 |
Tridecane | ||||||
Lipid14 | 6.47 | 3.53 | 1.84 | 0.81 | 0.82 | 0.83 |
Expt (46) | 6.5 | 3.5 | 1.86 | 0.68 | 0.64 | 0.77 |
Pentadecane | ||||||
Lipid14 | 7.80 | 4.20 | 1.86 | 0.81 | 1.00 | 1.02 |
Figure 6

Figure 6. Calculated 13C NMR T1 relaxation times for selected alkane chains and comparison to experiment. (47) Values at 312 K.
Lipid Bilayer Simulation
Initial Structures
no. of lipids | simulation time (ns) | temp (K) | waters/lipid nW | |
---|---|---|---|---|
DLPC | 128 | 5 × 125 | 303 | 31.3 |
DMPC | 128 | 5 × 125 | 303 | 25.6 |
DPPC | 128 | 5 × 125 | 323 | 30.1 |
DOPC | 128 | 5 × 125 | 303 | 32.8 |
POPC | 128 | 5 × 125 | 303 | 31 |
POPE | 128 | 5 × 125 | 310 | 32 |
Equilibration Procedure
Production Runs
Validation
Bilayer Structural Properties
lipid system | area per lipid AL (Å2) | volume per lipid VL (Å3) | isothermal area compressibility modulus KA (mNm-1) | bilayer thickness DHH (Å) | bilayer Luzzati thickness DB (Å) | ΔDB-H = (DB–DHH)/2 (Å) | ratio r of terminal methyl to methylene volume |
---|---|---|---|---|---|---|---|
DLPC | |||||||
Lipid14 | 63.0 ± 0.2 | 948.9 ± 0.3 | 281 ± 37 | 30.4 ± 0.4 | 30.2 ± 0.1 | –0.1 ± 0.2 | 1.9 |
Expt | 63.2, (54)60.8 (55) | 991 (54) | - | 30.8 (54) | 31.4 (54) | 0.8 (56) | 1.8–2.1 (57) |
DMPC | |||||||
Lipid14 | 59.7 ± 0.7 | 1050.2 ± 1.5 | 264 ± 90 | 34.7 ± 0.6 | 35.2 ± 0.4 | 0.3 ± 0.2 | 2.2 |
Expt | 60.6, (54)59.9 (55) | 1101 (4, 54) | 234 (58) | 34.4, (59) 35.3 (60) | 36.3, (54) 36.7, (55) | 0.8 (56) | 1.8–2.1 (57) |
36.9 (59) | |||||||
DPPC | |||||||
Lipid14 | 62.0 ± 0.3 | 1177.3 ± 0.5 | 244 ± 50 | 37.9 ± 0.5 | 38.0 ± 0.2 | 0.1 ± 0.2 | 2.1 |
Expt | 63.1, (55)64.3 (61) | 1232 (4) | 231 (4) | 38, (62) 38.3 (4) | 39.0 (55, 62) | 0.8 (56) | 1.8–2.1 (57) |
DOPC | |||||||
Lipid14 | 69.0 ± 0.3 | 1249.6 ± 0.2 | 338 ± 31 | 37.0 ± 0.2 | 36.2 ± 0.2 | –0.4 ± 0.1 | 2.1 |
Expt | 67.4, (62) 72.5 (4) | 1303 (4) | 265, (58) 300, (63) 318 (64) | 35.3, (65) 36.7, (62, 66) 36.9, (4) 37.1 (67) | 35.9, (4) 36.1, (65, 67) 38.7 (62) | 1.0–1.7 (57) | 1.8–2.1 (57) |
POPC | |||||||
Lipid14 | 65.6 ± 0.5 | 1205.4 ± 0.4 | 257 ± 47 | 36.9 ± 0.6 | 36.8 ± 0.3 | –0.1 ± 0.2 | 1.9 |
Expt | 64.3, (55) 68.3 (68) | 1256 (68) | 180–330 (69) | 37 (68) | 36.8, (68) 39.1 (55) | 0.8 (56) | 1.8–2.1 (57) |
POPE | |||||||
Lipid14 | 55.5 ± 0.2 | 1138.7 ± 0.3 | 350 ± 81 | 42.4 ± 0.2 | 41.0 ± 0.1 | –0.7 ± 0.1 | 2.0 |
Expt | 56.6, (70) 59–60 (71) | 1180 (71) | 233 (70) | 39.5 (71) | - | - | 1.8–2.1 (57) |
Ordering and Conformation of Lipid Acyl Chains

Figure 7

Figure 7. Simulation NMR order parameters for the six lipid systems and comparison to experiment. (77, 78, 80-84)
lipid system | eg | gg | gtg′ | gtg′+gtg | ng |
---|---|---|---|---|---|
DLPC | |||||
Lipid14 | 0.35 | 0.44 | 0.28 | 0.52 | 2.50 |
Expt (93) | 0.45 | 0.32 | 0.88* | - | 2.85 |
DMPC | |||||
Lipid14 | 0.34 | 0.48 | 0.35 | 0.62 | 2.82 |
Expt (94) | 0.38 | 0.67 | - | 0.44 | 2.6 |
DPPC | |||||
Lipid14 | 0.36 | 0.66 | 0.47 | 0.83 | 3.58 |
Expt | 0.38, (94) 0.4, (88) 0.54 (93) | 0.4, (88, 93) 0.57 (94) | 1.19 (93)a | 0.46, (94) 1.0 (88) | 2.44, (94) 3.6–4.2, (95) 3.7, (93) 3.8 (82) |
DOPC | |||||
Lipid14 | 0.36 | 0.75 | 0.37 | 0.70 | 3.93 |
POPC | |||||
Lipid14 | 0.36 | 0.69 | 0.41 | 0.75 | 3.73 |
POPE | |||||
Lipid14 | 0.35 | 0.60 | 0.42 | 0.73 | 3.50 |
Expt (88) | 0.05 | 0.2 | - | 0.8 | - |
The gtg′ sequence may be ascribed to a gtg′+gtg sequence. (90)
Electron Density Profiles
Figure 8

Figure 8. The total and decomposed electron density profiles for each of the six lipid bilayer systems with contributions from water, choline (CHOL), phosphate (PO4), glycerol (GLY), carbonyl (COO), methylene (CH2), unsaturated CH═CH and terminal methyls (CH3).
Scattering Form Factors
Figure 9

Figure 9. Simulation X-ray scattering form factors for the six lipid systems (black line) and comparison to experiment (54, 55, 62, 66, 68) (cyan circles). Inset: Simulation neutron scattering form factors at 100% D2O (black line), 70% D2O (red line), and 50% D2O (blue line) and comparison to experiment (55, 96) (black, red, and blue circles, respectively).
Figure 10

Figure 10. Plot of ΔDB-H versus area per lipid AL for the three all-atom lipid force fields CHARMM36 (squares), Slipids (diamonds), and AMBER Lipid14 (circles). Values shown for DLPC (green), DMPC (magenta), DPPC (blue), DOPC (red), and POPC (orange).
Lipid Lateral Diffusion
lipid system | calcd NPT Dxy (10-8 cm2 s-1) | calcd NVE Dxy (10-8 cm2 s-1) | simulation temp (K) | exptl Dxy (10-8 cm2 s-1) | exptl temp (K) |
---|---|---|---|---|---|
DLPC | 7.65 | 7.78 | 303 | 8.5 (97) | 298 |
DMPC | 5.05 | 6.32 | 303 | 5.95, (98) 9 (99, 100) | 303, 303 |
DPPC | 9.21 | 11.94 | 323 | 12.5, (101) 15.2 (102) | 323, 323 |
DOPC | 6.48 | 9.49 | 303 | 11.5, (100) 17 (103) | 303, 308 |
POPC | 5.74 | 6.54 | 303 | 10.7 (100) | 303 |
POPE | 4.67 | 4.85 | 310 | 5.2 (104)a | 305 |
Cell culture membrane containing 78% POPE at 305 K.
Figure 11

Figure 11. Time averaged mean square displacement of the center of mass of the lipid molecules versus NVE simulation time.
Figure 12

Figure 12. Lateral diffusion coefficients for the six lipid types calculated using different time ranges of the mean square displacement curve for the linear fit.
Conclusions
Supporting Information
Details of the Lipid14 atom types, partial charges, and force field parameters. Also included are the bilayer results for additional GPU and CPU runs. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
We are very grateful to Dr. Hannes Loeffler of the Science and Technology Facilities Council, UK, for writing and maintaining the modified PTRAJ/CPPTRAJ routines which were used for much of the analysis in this work. C.J.D. wishes to thank the Institute of Chemical Biology, UK Biotechnology and Biological Sciences Research Council (BBSRC) and GlaxoSmithKline for the award of a studentship, and the High Performance Computing centre of Imperial College London for the provision of computing time. B.D.M. would like to acknowledge funding for this work provided by the NIH Molecular Biophysics Training Grant (T32 GM008326) and the NVIDIA Graduate Fellowship Program. R.M.B. was supported by a grant from the University of California Institute for Mexico and the United States (UC MEXUS) and the Consejo Nacional de Ciencia y Tecnología de México (CONACYT) (R.C.W.). We acknowledge the support of the Strategic Programme for International Research and Education (SPIRE) and the Meltzer Foundation for travel grants provided to Å.A.S. The Norwegian Metacenter for Computational Science (NOTUR) is acknowledged for allocation of computational resources. This work was supported by NSF SI2-SSE grants (NSF-1047875 and 1148276) to R.C.W. and by the University of California (UC Lab 09-LR-06-117793) grant to R.C.W. R.C.W. also acknowledges funding through the NSF XSEDE program and through a fellowship from NVIDIA, Inc. Additional computer time was provided by the San Diego Supercomputer Center and by XSEDE and TG-CHG13W10 to R.C.W.
References
This article references 109 other publications.
- 1van Meer, G.; Voelker, D. R.; Feigenson, G. W. Membrane lipids: where they are and how they behave Nat. Rev. Mol. Cell Biol. 2008, 9 (2) 112– 124[Crossref], [PubMed], [CAS], Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXovFOntw%253D%253D&md5=13d4eb41956211e22d3a9ee5c6b73854Membrane lipids: where they are and how they behavevan Meer, Gerrit; Voelker, Dennis R.; Feigenson, Gerald W.Nature Reviews Molecular Cell Biology (2008), 9 (2), 112-124CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. Throughout the biol. world, a 30 Å hydrophobic film typically delimits the environments that serve as the margin between life and death for individual cells. Biochem. and biophys. findings have provided a detailed model of the compn. and structure of membranes, which includes levels of dynamic organization both across the lipid bilayer (lipid asymmetry) and in the lateral dimension (lipid domains) of membranes. How do cells apply anabolic and catabolic enzymes, translocases, and transporters, plus the intrinsic phys. phase behavior of lipids and their interactions with membrane proteins, to create the unique compns. and multiple functionalities of their individual membranes.
- 2Lodish, H.; Berk, A.; Kaiser, C. A.; Scott, M. P.; Bretscher, A.; Ploegh, H.; Matsudaira, P. Molecular Cell Biology,6th ed.; W. H. Freeman: New York, 2007.Google ScholarThere is no corresponding record for this reference.
- 3Phillips, R.; Ursell, T.; Wiggins, P.; Sens, P. Emerging roles for lipids in shaping membrane-protein function Nature 2009, 459 (7245) 379– 385[Crossref], [PubMed], [CAS], Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmtFKmtrc%253D&md5=802acc0d4cf9ce32bd821d46f736e1ccEmerging roles for lipids in shaping membrane-protein functionPhillips, Rob; Ursell, Tristan; Wiggins, Paul; Sens, PierreNature (London, United Kingdom) (2009), 459 (7245), 379-385CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. Studies of membrane proteins have revealed a direct link between the lipid environment and the structure and function of some of these proteins. Although some of these effects involve specific chem. interactions between lipids and protein residues, many can be understood in terms of protein-induced perturbations to the membrane shape. The free-energy cost of such perturbations can be estd. quant., and measurements of channel gating in model systems of membrane proteins with their lipid partners are now confirming predictions of simple models.
- 4Nagle, J. F.; Tristram-Nagle, S. Structure of lipid bilayers Biochim. Biophys. Acta 2000, 1469 (3) 159– 195
- 5Katsaras, J.; Gutberlet, T. Lipid bilayers: Structure and interactions; Springer-Verlag: Berlin, 2001.
- 6Tieleman, D. P.; Marrink, S. J.; Berendsen, H. J. A computer perspective of membranes: molecular dynamics studies of lipid bilayer systems Biochim. Biophys. Acta 1997, 1331, 235– 270[Crossref], [PubMed], [CAS], Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXnvVSnsLw%253D&md5=58cecf9cb3f6c4d20d5a717eefbc4698A computer perspective of membranes: molecular dynamics studies of lipid bilayer systemsTieleman, D. P.; Marrink, S. J.; Berendsen, H. J. C.Biochimica et Biophysica Acta, Reviews on Biomembranes (1997), 1331 (3), 235-270CODEN: BRBMC5; ISSN:0304-4157. (Elsevier B.V.)A review with 125 refs. on mol. dynamic simulations and exptl. data which can be used to validate the simulations of lipid bilayers.
- 7Berger, O.; Edholm, O.; Jähnig, F. Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperature Biophys. J. 1997, 72, 2002– 2013[Crossref], [PubMed], [CAS], Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXivVOjsL4%253D&md5=984ea49ced9d0e2ce912bf99220fd228Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperatureBerger, Oliver; Edholm, Olle; Jahnig, FritzBiophysical Journal (1997), 72 (5), 2002-2013CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Mol. dynamics simulations of 500 ps were performed on a system consisting of a bilayer of 64 mols. of the lipid dipalmitoylphosphatidylcholine and 23 water mols. per lipid at an isotropic pressure of 1 atm and 50°. Special attention was devoted to reproduce the correct d. of the lipid, because this quantity is known exptl. with a precision better than 1%. For this purpose, the Lennard-Jones parameters of the hydrocarbon chains were adjusted by simulating a system consisting of 128 pentadecane mols. and varying the Lennard-Jones parameters until the exptl. d. and heat of vaporization were obtained. With these parameters the lipid d. resulted in perfect agreement with the exptl. d. The orientational order parameter of the hydrocarbon chains agreed perfectly well with the exptl. values, which, because of its correlation with the area per lipid, makes it possible to give a proper est. of the area per lipid of 0.61±0.1 nm2.
- 8Klauda, J. B.; Venable, R. M.; Freites, J. A.; O’Connor, J. W.; Tobias, D. J.; Mondragon-Ramirez, C.; Vorobyov, I.; MacKerell, A. D.; Pastor, R. W. Update of the CHARMM all-atom additive force field for lipids: Validation on Six lipid types J. Phys. Chem. B 2010, 114 (23) 7830– 7843[ACS Full Text
], [CAS], Google Scholar
8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXmsVGjtrk%253D&md5=04c4d3aa1a59740190994d7df80cbb71Update of the CHARMM All-Atom Additive Force Field for Lipids: Validation on Six Lipid TypesKlauda, Jeffery B.; Venable, Richard M.; Freites, J. Alfredo; O'Connor, Joseph W.; Tobias, Douglas J.; Mondragon-Ramirez, Carlos; Vorobyov, Igor; MacKerell, Alexander D., Jr.; Pastor, Richard W.Journal of Physical Chemistry B (2010), 114 (23), 7830-7843CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)A significant modification to the additive all-atom CHARMM lipid force field (FF) is developed and applied to phospholipid bilayers with both choline and ethanolamine contg. head groups and with both satd. and unsatd. aliph. chains. Motivated by the current CHARMM lipid FF (C27 and C27r) systematically yielding values of the surface area per lipid that are smaller than exptl. ests. and gel-like structures of bilayers well above the gel transition temp., selected torsional, Lennard-Jones and partial at. charge parameters were modified by targeting both quantum mech. (QM) and exptl. data. QM calcns. ranging from high-level ab initio calcns. on small mols. to semiempirical QM studies on a 1,2-dipalmitoyl-sn-phosphatidylcholine (DPPC) bilayer in combination with exptl. thermodn. data were used as target data for parameter optimization. These changes were tested with simulations of pure bilayers at high hydration of the following six lipids: DPPC, 1,2-dimyristoyl-sn-phosphatidylcholine (DMPC), 1,2-dilauroyl-sn-phosphatidylcholine (DLPC), 1-palmitoyl-2-oleoyl-sn-phosphatidylcholine (POPC), 1,2-dioleoyl-sn-phosphatidylcholine (DOPC), and 1-palmitoyl-2-oleoyl-sn-phosphatidylethanolamine (POPE); simulations of a low hydration DOPC bilayer were also performed. Agreement with exptl. surface area is on av. within 2%, and the d. profiles agree well with neutron and x-ray diffraction expts. NMR deuterium order parameters (SCD) are well predicted with the new FF, including proper splitting of the SCD for the aliph. carbon adjacent to the carbonyl for DPPC, POPE, and POPC bilayers. The area compressibility modulus and frequency dependence of 13C NMR relaxation rates of DPPC and the water distribution of low hydration DOPC bilayers also agree well with expt. Accordingly, the presented lipid FF, referred to as C36, allows for mol. dynamics simulations to be run in the tensionless ensemble (NPT), and is anticipated to be of utility for simulations of pure lipid systems as well as heterogeneous systems including membrane proteins. - 9Poger, D.; Van Gunsteren, W. F.; Mark, A. E. A new force field for simulating phosphatidylcholine bilayers J. Comput. Chem. 2010, 31 (6) 1117– 1125[Crossref], [PubMed], [CAS], Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjsFSkt7c%253D&md5=a1626ceb93b19b562b47a7c7a7d36b44A new force field for simulating phosphatidylcholine bilayersPoger, David; Van Gunsteren, Wilfred F.; Mark, Alan E.Journal of Computational Chemistry (2010), 31 (6), 1117-1125CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new force field for the simulation of dipalmitoylphosphatidylcholine (DPPC) in the liq.-cryst., fluid phase at zero surface tension is presented. The structure of the bilayer with the area per lipid (0.629 Nm2; expt. 0.629-0.64 Nm2), the vol. per lipid (1.226 Nm3; expt. 1.229-1.232 Nm3), and the ordering of the palmitoyl chains (order parameters) are all in very good agreement with expt. Exptl. electron d. profiles are well reproduced in particular with regard to the penetration of water into the bilayer. The force field was further validated by simulating the spontaneous assembly of DPPC into a bilayer in water. Notably, the timescale on which membrane sealing was obsd. using this model appears closer to the timescales for membrane resealing suggested by electroporation expts. than previous simulations using existing models. © 2009 Wiley Periodicals, Inc.
- 10Jämbeck, J. P. M.; Lyubartsev, A. P. Derivation and systematic validation of a refined all-atom force field for phosphatidylcholine lipids J. Phys. Chem. B 2012, 116 (10) 3164– 3179
- 11Marrink, S. J.; Risselada, H. J.; Yefimov, S.; Tieleman, D. P.; de Vries, A. H. The MARTINI force field: Coarse grained model for biomolecular simulations J. Phys. Chem. B 2007, 111 (27) 7812– 7824[ACS Full Text
], [CAS], Google Scholar
11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXmsVKmsLc%253D&md5=428d5750b94652e4917d905a30658235The MARTINI Force Field: Coarse Grained Model for Biomolecular SimulationsMarrink, Siewert J.; Risselada, H. Jelger; Yefimov, Serge; Tieleman, D. Peter; De Vries, Alex H.Journal of Physical Chemistry B (2007), 111 (27), 7812-7824CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present an improved and extended version of our coarse grained lipid model. The new version, coined the MARTINI force field, is parametrized in a systematic way, based on the reprodn. of partitioning free energies between polar and apolar phases of a large no. of chem. compds. To reproduce the free energies of these chem. building blocks, the no. of possible interaction levels of the coarse-grained sites has increased compared to those of the previous model. Application of the new model to lipid bilayers shows an improved behavior in terms of the stress profile across the bilayer and the tendency to form pores. An extension of the force field now also allows the simulation of planar (ring) compds., including sterols. Application to a bilayer/cholesterol system at various concns. shows the typical cholesterol condensation effect similar to that obsd. in all atom representations. - 12Orsi, M.; Essex, J. W. The ELBA force field for coarse-grain modeling of lipid membranes PLoS One 2011, 6 (12) e28637[Crossref], [PubMed], [CAS], Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjvFahsA%253D%253D&md5=feb2d1452c92ecce6a91dc2d28496690The ELBA force field for coarse-grain modeling of lipid membranesOrsi, Mario; Essex, Jonathan W.PLoS One (2011), 6 (12), e28637CODEN: POLNCL; ISSN:1932-6203. (Public Library of Science)A new coarse-grain model for mol. dynamics simulation of lipid membranes is presented. Following a simple and conventional approach, lipid mols. are modeled by spherical sites, each representing a group of several atoms. In contrast to common coarse-grain methods, two original (interdependent) features are here adopted. First, the main electrostatics are modeled explicitly by charges and dipoles, which interact realistically through a relative dielec. const. of unity (εr = 1). Second, water mols. are represented individually through a new parametrization of the simple Stockmayer potential for polar fluids; each water mol. is therefore described by a single spherical site embedded with a point dipole. The force field is shown to accurately reproduce the main phys. properties of single-species phospholipid bilayers comprising dioleoylphosphatidylcholine (DOPC) and dioleoylphosphatidylethanolamine (DOPE) in the liq. crystal phase, as well as distearoylphosphatidylcholine (DSPC) in the liq. crystal and gel phases. Insights are presented into fundamental properties and phenomena that can be difficult or impossible to study with alternative computational or exptl. methods. For example, we investigate the internal pressure distribution, dipole potential, lipid diffusion, and spontaneous self-assembly. Simulations lasting up to 1.5 μs were conducted for systems of different sizes (128, 512 and 1058 lipids); this also allowed us to identify size-dependent artifacts that are expected to affect membrane simulations in general. Future extensions and applications are discussed, particularly in relation to the methodol.'s inherent multiscale capabilities.
- 13Chiu, S.-W.; Pandit, S. A.; Scott, H. L.; Jakobsson, E. An improved united atom force field for simulation of mixed lipid bilayers J. Phys. Chem. B 2009, 113 (9) 2748– 2763[ACS Full Text
], [CAS], Google Scholar
13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Sltro%253D&md5=c881cacab451368c125262e4b9dd0e01An Improved United Atom Force Field for Simulation of Mixed Lipid BilayersChiu, See-Wing; Pandit, Sagar A.; Scott, H. L.; Jakobsson, EricJournal of Physical Chemistry B (2009), 113 (9), 2748-2763CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We introduce a new force field (43A1-S3) for simulation of membranes by the Gromacs simulation package. Construction of the force fields is by std. methods of electronic structure computations for bond parameters and charge distribution and sp. vol.s and heats of vaporization for small-mol. components of the larger lipid mols. for van der Waals parameters. Some parameters from the earlier 43A1 force field are found to be correct in the context of these calcns., while others are modified. The validity of the force fields is demonstrated by correct replication of X-ray form factors and NMR order parameters over a wide range of membrane compns. in semi-isotropic NTP 1 atm simulations. 43-A1-S3 compares favorably with other force fields used in conjunction with the Gromacs simulation package with respect to the breadth of phenomena that it accurately reproduces. - 14Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of multiple Amber force fields and development of improved protein backbone parameters Proteins: Struct., Funct., Bioinf. 2006, 65 (3) 712– 725
- 15Kirschner, K. N.; Yongye, A. B.; Tschampel, S. M.; González-Outeiriño, J.; Daniels, C. R.; Foley, B. L.; Woods, R. J. GLYCAM06: A generalizable biomolecular force field. Carbohydrates J. Comput. Chem. 2008, 29 (4) 622– 655
- 16Wang, 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, 1157– 1174[Crossref], [PubMed], [CAS], Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development 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.
- 17Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Swails, J.; Goetz, A. W.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Wolf, R. M.; Liu, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M.-J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Salomon-Ferrer, R.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A.AMBER 12; University of California: San Francisco, 2012.Google ScholarThere is no corresponding record for this reference.
- 18Salomon-Ferrer, R.; Case, D. A.; Walker, R. C. An overview of the Amber biomolecular simulation package Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2013, 3 (2) 198– 210[Crossref], [CAS], Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXmtFaiu70%253D&md5=baac68008e08e80bd825b2362e14fa1fAn overview of the amber biomolecular simulation packageSalomon-Ferrer, Romelia; Case, David A.; Walker, Ross C.Wiley Interdisciplinary Reviews: Computational Molecular Science (2013), 3 (2), 198-210CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Mol. dynamics (MD) allows the study of biol. and chem. systems at the atomistic level on timescales from femtoseconds to milliseconds. It complements expt. while also offering a way to follow processes difficult to discern with exptl. techniques. Numerous software packages exist for conducting MD simulations of which one of the widest used is termed Amber. Here, we outline the most recent developments, since version 9 was released in Apr. 2006, of the Amber and AmberTools MD software packages, referred to here as simply the Amber package. The latest release represents six years of continued development, since version 9, by multiple research groups and the culmination of over 33 years of work beginning with the first version in 1979. The latest release of the Amber package, version 12 released in Apr. 2012, includes a substantial no. of important developments in both the scientific and computer science arenas. We present here a condensed vision of what Amber currently supports and where things are likely to head over the coming years. Figure 1 shows the performance in ns/day of the Amber package version 12 on a single-core AMD FX-8120 8-Core 3.6GHz CPU, the Cray XT5 system, and a single GPU GTX680.
- 19Götz, A. W.; Williamson, M. J.; Xu, D.; Poole, D.; Le Grand, S.; Walker, R. C. Routine microsecond molecular dynamics simulations with AMBER on GPUs. 1. Generalized Born J. Chem. Theory Comput. 2012, 8 (5) 1542– 1555[ACS Full Text
], [CAS], Google Scholar
19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XksFWns78%253D&md5=1e6f570db9cd504bb13706e7c56bc356Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 1. Generalized BornGotz, Andreas W.; Williamson, Mark J.; Xu, Dong; Poole, Duncan; Le Grand, Scott; Walker, Ross C.Journal of Chemical Theory and Computation (2012), 8 (5), 1542-1555CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present an implementation of generalized Born implicit solvent all-atom classical mol. dynamics (MD) within the AMBER program package that runs entirely on CUDA enabled NVIDIA graphics processing units (GPUs). We discuss the algorithms that are used to exploit the processing power of the GPUs and show the performance that can be achieved in comparison to simulations on conventional CPU clusters. The implementation supports three different precision models in which the contributions to the forces are calcd. in single precision floating point arithmetic but accumulated in double precision (SPDP), or everything is computed in single precision (SPSP) or double precision (DPDP). In addn. to performance, we have focused on understanding the implications of the different precision models on the outcome of implicit solvent MD simulations. We show results for a range of tests including the accuracy of single point force evaluations and energy conservation as well as structural properties pertaining to protein dynamics. The numerical noise due to rounding errors within the SPSP precision model is sufficiently large to lead to an accumulation of errors which can result in unphys. trajectories for long time scale simulations. We recommend the use of the mixed-precision SPDP model since the numerical results obtained are comparable with those of the full double precision DPDP model and the ref. double precision CPU implementation but at significantly reduced computational cost. Our implementation provides performance for GB simulations on a single desktop that is on par with, and in some cases exceeds, that of traditional supercomputers. - 20Salomon-Ferrer, R.; Götz, A. W.; Poole, D.; Le Grand, S.; Walker, R. C. Routine microsecond molecular dynamics simulations with Amber on GPUs. 2. Explicit solvent particle mesh Ewald J. Chem. Theory Comput. 2013, 9 (9) 3878– 3888[ACS Full Text
], [CAS], Google Scholar
20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1arsrzP&md5=57e18d35e8b81e9a79788ef5af17d19fRoutine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh EwaldSalomon-Ferrer, Romelia; Gotz, Andreas W.; Poole, Duncan; Le Grand, Scott; Walker, Ross C.Journal of Chemical Theory and Computation (2013), 9 (9), 3878-3888CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present an implementation of explicit solvent all atom classical mol. dynamics (MD) within the AMBER program package that runs entirely on CUDA-enabled GPUs. First released publicly in Apr. 2010 as part of version 11 of the AMBER MD package and further improved and optimized over the last two years, this implementation supports the three most widely used statistical mech. ensembles (NVE, NVT, and NPT), uses particle mesh Ewald (PME) for the long-range electrostatics, and runs entirely on CUDA-enabled NVIDIA graphics processing units (GPUs), providing results that are statistically indistinguishable from the traditional CPU version of the software and with performance that exceeds that achievable by the CPU version of AMBER software running on all conventional CPU-based clusters and supercomputers. We briefly discuss three different precision models developed specifically for this work (SPDP, SPFP, and DPDP) and highlight the tech. details of the approach as it extends beyond previously reported work [Goetz et al., J. Chem. Theory Comput.2012, DOI: 10.1021/ct200909j; Le Grand et al., Comp. Phys. Comm.2013, DOI: 10.1016/j.cpc.2012.09.022].We highlight the substantial improvements in performance that are seen over traditional CPU-only machines and provide validation of our implementation and precision models. We also provide evidence supporting our decision to deprecate the previously described fully single precision (SPSP) model from the latest release of the AMBER software package. - 21Le Grand, S.; Götz, A. W.; Walker, R. C. SPFP: Speed without compromise—A mixed precision model for GPU accelerated molecular dynamics simulations Comput. Phys. Commun. 2013, 184 (2) 374– 380[Crossref], [CAS], Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsV2jsrzO&md5=2d9f102cb17f6e9bcbff3b5057d8c587SPFP: Speed without compromise-A mixed precision model for GPU accelerated molecular dynamics simulationsLe Grand, Scott; Gotz, Andreas W.; Walker, Ross C.Computer Physics Communications (2013), 184 (2), 374-380CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier B.V.)A new precision model is proposed for the acceleration of all-atom classical mol. dynamics (MD) simulations on graphics processing units (GPUs). This precision model replaces double precision arithmetic with fixed point integer arithmetic for the accumulation of force components as compared to a previously introduced model that uses mixed single/double precision arithmetic. This significantly boosts performance on modern GPU hardware without sacrificing numerical accuracy. We present an implementation for NVIDIA GPUs of both generalized Born implicit solvent simulations as well as explicit solvent simulations using the particle mesh Ewald (PME) algorithm for long-range electrostatics using this precision model. Tests demonstrate both the performance of this implementation as well as its numerical stability for const. energy and const. temp. biomol. MD as compared to a double precision CPU implementation and double and mixed single/double precision GPU implementations.
- 22Skjevik, Å. A.; Madej, B. D.; Walker, R. C.; Teigen, K. LIPID11: A modular framework for lipid simulations using Amber J. Phys. Chem. B 2012, 116 (36) 11124– 11136
- 23Siu, S. W.; Vacha, R.; Jungwirth, P.; Bockmann, R. A. Biomolecular simulations of membranes: physical properties from different force fields J. Chem. Phys. 2008, 128 (12) 125103[Crossref], [PubMed], [CAS], Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXktl2nsro%253D&md5=38ee4310c79b26dac45f24a14aa7d8e7Biomolecular simulations of membranes: Physical properties from different force fieldsSiu, Shirley W. I.; Vacha, Robert; Jungwirth, Pavel; Boeckmann, Rainer A.Journal of Chemical Physics (2008), 128 (12), 125103/1-125103/12CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Phospholipid force fields are of ample importance for the simulation of artificial bilayers, membranes, and also for the simulation of integral membrane proteins. Here, we compare the two most applied at. force fields for phospholipids, the all-atom CHARMM27 and the united atom Berger force field, with a newly developed all-atom generalized AMBER force field (GAFF) for dioleoylphosphatidylcholine mols. Only the latter displays the exptl. obsd. difference in the order of the C2 atom between the two acyl chains. The interfacial water dynamics is smoothly increased between the lipid carbonyl region and the bulk water phase for all force fields; however, the water order and with it the electrostatic potential across the bilayer showed distinct differences between the force fields. Both Berger and GAFF underestimate the lipid self-diffusion. GAFF offers a consistent force field for the at. scale simulation of biomembranes. (c) 2008 American Institute of Physics.
- 24Jójárt, B.; Martinek, T. A. Performance of the general amber force field in modeling aqueous POPC membrane bilayers J. Comput. Chem. 2007, 28 (12) 2051– 2058
- 25Rosso, L.; Gould, I. R. Structure and dynamics of phospholipid bilayers using recently developed general all-atom force fields J. Comput. Chem. 2008, 29 (1) 24– 37
- 26Dickson, C. J.; Rosso, L.; Betz, R. M.; Walker, R. C.; Gould, I. R. GAFFlipid: A general Amber force field for the accurate molecular dynamics simulation of phospholipid Soft Matter 2012, 8, 9617– 9627[Crossref], [CAS], Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1ymsbzL&md5=494ab83d6e79d084d2876c2169b9add1GAFFlipid: a General Amber Force Field for the accurate molecular dynamics simulation of phospholipidDickson, Callum J.; Rosso, Lula; Betz, Robin M.; Walker, Ross C.; Gould, Ian R.Soft Matter (2012), 8 (37), 9617-9627CODEN: SMOABF; ISSN:1744-683X. (Royal Society of Chemistry)Previous attempts to simulate phospholipid bilayers using the General Amber Force Field (GAFF) yielded many bilayer characteristics in agreement with expt., however when using a tensionless NPT ensemble the bilayer is seen to compress to an undesirable extent resulting in low areas per lipid and high order parameters in comparison to expt. In this work, the GAFF Lennard-Jones parameters for the simulation of acyl chains are cor. to allow the accurate and stable simulation of pure lipid bilayers. Lipid bilayers comprised of six phospholipid types were simulated for timescales approaching a quarter of a microsecond under tensionless const. pressure conditions using Graphics Processing Units. Structural properties including area per lipid, vol. per lipid, bilayer thickness, order parameter and headgroup hydration show favorable agreement with available exptl. values. Expanding the system size from 72 to 288 lipids and a more exptl. realistic 2 × 288 lipid bilayer stack induces little change in the obsd. properties. This preliminary work is intended for combination with the new AMBER Lipid11 modular force field as part of on-going attempts to create a modular phospholipid AMBER force field allowing tensionless NPT simulations of complex lipid bilayers.
- 27Siu, S. W. I.; Pluhackova, K.; Böckmann, R. A. Optimization of the OPLS-AA force field for long hydrocarbons J. Chem. Theory Comput. 2012, 8 (4) 1459– 1470[ACS Full Text
], [CAS], Google Scholar
27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjsFeiu78%253D&md5=aa2ce129ad72298695a988fe0e37af7dOptimization of the OPLS-AA Force Field for Long HydrocarbonsSiu, Shirley W. I.; Pluhackova, Kristyna; Boeckmann, Rainer A.Journal of Chemical Theory and Computation (2012), 8 (4), 1459-1470CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The all-atom optimized potentials for liq. simulations (OPLS-AA) force field is a popular force field for simulating biomols. However, the current OPLS parameters for hydrocarbons developed using short alkanes cannot reproduce the liq. properties of long alkanes in mol. dynamics simulations. Therefore, the extension of OPLS-AA to (phospho)lipid mols. required for the study of biol. membranes was hampered in the past. Here, we optimized the OPLS-AA force field for both short and long hydrocarbons. Following the framework of the OPLS-AA parametrization, we refined the torsional parameters for hydrocarbons by fitting to the gas-phase ab initio energy profiles calcd. at the accurate MP2/aug-cc-pVTZ theory level. Addnl., the depth of the Lennard-Jones potential for methylene hydrogen atoms was adjusted to reproduce the densities and the heats of vaporization of alkanes and alkenes of different lengths. Optimization of partial charges finally allowed to reproduce the gel-to-liq.-phase transition temp. for pentadecane and solvation free energies. The optimized parameter set (L-OPLS) yields improved hydrocarbon diffusion coeffs., viscosities, and gauche-trans ratios. Moreover, its applicability for lipid bilayer simulations is shown for a GMO bilayer in its liq.-cryst. phase. - 28Klauda, J. B.; Brooks, B. R.; MacKerell, A. D.; Venable, R. M.; Pastor, R. W. An ab initio study on the torsional surface of alkanes and its effect on molecular simulations of alkanes and a DPPC bilayer J. Phys. Chem. B 2005, 109 (11) 5300– 5311
- 29Betz, R. M.; Walker, R. C. Paramfit: Optimization of potential energy function parameters for molecular dynamics. Manuscript in preparation.Google ScholarThere is no corresponding record for this reference.
- 30Bayly, C. I.; Cieplak, P.; Cornell, W.; Kollman, P. A. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model J. Phys. Chem. 1993, 97 (40) 10269– 10280[ACS Full Text
], [CAS], Google Scholar
30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXlvVyqsLs%253D&md5=e65c6a556ffc174df4f327687912a0bdA well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP modelBayly, Christopher I.; Cieplak, Piotr; Cornell, Wendy; Kollman, Peter A.Journal of Physical Chemistry (1993), 97 (40), 10269-80CODEN: JPCHAX; ISSN:0022-3654.The authors present a new approach to generating electrostatic potential (ESP) derived charges for mols. The major strength of electrostatic potential derived charges is that they optimally reproduce the intermol. interaction properties of mols. with a simple two-body additive potential, provided, of course, that a suitably accurate level of quantum mech. calcn. is used to derive the ESP around the mol. Previously, the major weaknesses of these charges have been that they were not easily transferably between common functional groups in related mols., they have often been conformationally dependent, and the large charges that frequently occur can be problematic for simulating intramol. interactions. Introducing restraints in the form of a penalty function into the fitting process considerably reduces the above problems, with only a minor decrease in the quality of the fit to the quantum mech. ESP. Several other refinements in addn. to the restrained electrostatic potential (RESP) fit yield a general and algorithmic charge fitting procedure for generating atom-centered point charges. This approach can thus be recommended for general use in mol. mechanics, mol. dynamics, and free energy calcns. for any org. or bioorg. system. - 31Sonne, J.; Jensen, M. Ø.; Hansen, F. Y.; Hemmingsen, L.; Peters, G. H. Reparameterization of all-atom dipalmitoylphosphatidylcholine lipid parameters enables simulation of fluid bilayers at zero tension Biophys. J. 2007, 92 (12) 4157– 4167
- 32Klauda, J. B.; Garrison, S. L.; Jiang, J.; Arora, G.; Sandler, S. I. HM-IE: Quantum chemical hybrid methods for calculating interaction energies J. Phys. Chem. A 2003, 108 (1) 107– 112
- 33Davis, J. E.; Warren, G. L.; Patel, S. Revised charge equilibration potential for liquid alkanes J. Phys. Chem. B 2008, 112 (28) 8298– 8310
- 34Wang, J.; Hou, T. Application of molecular dynamics simulations in molecular property prediction. 1. density and heat of vaporization J. Chem. Theory Comput. 2011, 7 (7) 2151– 2165[ACS Full Text
], [CAS], Google Scholar
34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnt1Gmtbw%253D&md5=3485eec18e247b268e9733d189465f02Application of Molecular Dynamics Simulations in Molecular Property Prediction. 1. Density and Heat of VaporizationWang, Junmei; Hou, TingjunJournal of Chemical Theory and Computation (2011), 7 (7), 2151-2165CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Mol. mech. force field (FF) methods are useful in studying condensed phase properties. They are complementary to expts. and can often go beyond expts. in at. details. Even if a FF is specific for studying structures, dynamics, and functions of biomols., it is still important for the FF to accurately reproduce the exptl. liq. properties of small mols. that represent the chem. moieties of biomols. Otherwise, the force field may not describe the structures and energies of macromols. in aq. solns. properly. In this work, the authors have carried out a systematic study to evaluate the General AMBER Force Field (GAFF) in studying densities and heats of vaporization for a large set of org. mols. that covers the most common chem. functional groups. The latest techniques, such as the particle mesh Ewald (PME) for calcg. electrostatic energies and Langevin dynamics for scaling temps., have been applied in the mol. dynamics (MD) simulations. For d., the av. percent error (APE) of 71 org. compds. is 4.43% when compared with the exptl. values. More encouragingly, the APE drops to 3.43% after the exclusion of two outliers and four other compds. for which the exptl. densities have been measured with pressures higher than 1.0 atm. For the heat of vaporization, several protocols have been investigated, and the best one, P4/ntt0, achieves an av. unsigned error (AUE) and a root-mean-square error (RMSE) of 0.93 and 1.20 kcal/mol, resp. How to reduce the prediction errors through proper van der Waals (vdW) parametrization has been discussed. An encouraging finding in vdW parametrization is that both densities and heats of vaporization approach their "ideal" values in a synchronous fashion when vdW parameters are tuned. The following hydration free energy calcn. using thermodn. integration further justifies the vdW refinement. The authors conclude that simple vdW parametrization can significantly reduce the prediction errors. The authors believe that GAFF can greatly improve its performance in predicting liq. properties of org. mols. after a systematic vdW parametrization, which will be reported in a sep. paper. - 35Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N-log(N) method for Ewald sums in large systems J. Chem. Phys. 1993, 98 (12) 10089– 10092
- 36Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. C. Numerical integration of the cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes J. Comput. Phys. 1977, 23 (3) 327– 341[Crossref], [CAS], Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2sXktVGhsL4%253D&md5=b4aecddfde149117813a5ea4f5353ce2Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanesRyckaert, Jean Paul; Ciccotti, Giovanni; Berendsen, Herman J. C.Journal of Computational Physics (1977), 23 (3), 327-41CODEN: JCTPAH; ISSN:0021-9991.A numerical algorithm integrating the 3N Cartesian equation of motion of a system of N points subject to holonomic constraints is applied to mol. dynamics simulation of a liq. of 64 butane mols.
- 37Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R. Molecular dynamics with coupling to an external bath J. Chem. Phys. 1984, 81 (8) 3684– 3690[Crossref], [CAS], Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXmtlGksbY%253D&md5=5510dc00297d63b91ee3a7a4ae5aacb1Molecular dynamics with coupling to an external bathBerendsen, H. J. C.; Postma, J. P. M.; Van Gunsteren, W. F.; DiNola, A.; Haak, J. R.Journal of Chemical Physics (1984), 81 (8), 3684-90CODEN: JCPSA6; ISSN:0021-9606.In mol. dynamics (MD) simulations, the need often arises to maintain such parameters as temp. or pressure rather than energy and vol., or to impose gradients for studying transport properties in nonequil. MD. A method is described to realize coupling to an external bath with const. temp. or pressure with adjustable time consts. for the coupling. The method is easily extendable to other variables and to gradients, and can be applied also to polyat. mols. involving internal constraints. The influence of coupling time consts. on dynamical variables is evaluated. A leap-frog algorithm is presented for the general case involving constraints with coupling to both a const. temp. and a const. pressure bath.
- 38Yeh, I.-C.; Hummer, G. System-size dependence of diffusion coefficients and viscosities from molecular dynamics simulations with periodic boundary conditions J. Phys. Chem. B 2004, 108 (40) 15873– 15879[ACS Full Text
], [CAS], Google Scholar
38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXnsV2iurk%253D&md5=e333a4ea27b4451db997737f0e3f6a4fSystem-Size Dependence of Diffusion Coefficients and Viscosities from Molecular Dynamics Simulations with Periodic Boundary ConditionsYeh, In-Chul; Hummer, GerhardJournal of Physical Chemistry B (2004), 108 (40), 15873-15879CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We study the system-size dependence of translational diffusion coeffs. and viscosities in mol. dynamics simulations under periodic boundary conditions. Simulations of water under ambient conditions and a Lennard-Jones (LJ) fluid show that the diffusion coeffs. increase strongly as the system size increases. We test a simple analytic correction for the system-size effects that is based on hydrodynamic arguments. This correction scales as N-1/3, where N is the no. of particles. For a cubic simulation box of length L, the diffusion coeff. cor. for system-size effects is D0 = DPBC + 2.837297kBT/(6πηL), where DPBC is the diffusion coeff. calcd. in the simulation, kB the Boltzmann const., T the abs. temp., and η the shear viscosity of the solvent. For water, LJ fluids, and hard-sphere fluids, this correction quant. accounts for the system-size dependence of the calcd. self-diffusion coeffs. In contrast to diffusion coeffs., the shear viscosities of water and the LJ fluid show no significant system-size dependences. - 39Lipari, G.; Szabo, A. Effect of librational motion on fluorescence depolarization and nuclear magnetic resonance relaxation in macromolecules and membranes Biophys. J. 1980, 30 (3) 489– 506[Crossref], [PubMed], [CAS], Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXltV2gtLo%253D&md5=066330f119aa1ccabb117d13f54b02c4Effect of librational motion on fluorescence depolarization and nuclear magnetic resonance relaxation in macromolecules and membranesLipari, Giovanni; Szabo, AttilaBiophysical Journal (1980), 30 (3), 489-506CODEN: BIOJAU; ISSN:0006-3495.The theory of fluorescent emission anisotropy (r(t)) of a cylindrical probe in a membrane suspension is developed. The limiting anisotropy (r(∞)) is proportional to the square of the order parameter of the probe. The order parameter dets. the 1st nontrivial term in the expansion of the equil. orientational distribution function of the probe in a series of Legendre polynomials. The motion of the probe is described as diffusion (wobbling) within a cone of semiangle θ0. Within the framework of this model, an accurate single-exponential approxn. for r(t) is considered. An analytic expression relating the effective relaxation time, which appears in the above approxn., to θ0 and the diffusion coeff. for wobbling is derived. The model is generalized to the situation where the probe is attached to a macromol. whose motion cannot be neglected on the time scale of the fluorescence expt. Finally, by exploiting the formal similarity between the theory of fluorescence depolarization and 13C NMR dipolar relaxation, expressions for spin-spin and spin-lattice relaxation times and the nuclear Overhauser enhancement are derived for a protonated C which is nonrigidly attached to a macromol. and undergoes librational motion described as diffusion on a spherical cap of semiangle θ0.
- 40Ottiger, M.; Bax, A. Determination of Relative N–HN, N–C′, Cα–C′, and Cα–Hα effective bond lengths in a protein by NMR in a dilute liquid crystalline phase J. Am. Chem. Soc. 1998, 120 (47) 12334– 12341
- 41Haynes, W. M. CRC Handbook of Chemistry and Physics, 93rd ed.; CRC Press: Boca Raton, FL, 2012–2013.Google ScholarThere is no corresponding record for this reference.
- 42Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.Gaussian 09, Revision A.1; Gaussian Inc.: Wallingford, CT, 2009.Google ScholarThere is no corresponding record for this reference.
- 43Douglass, D. C.; McCall, D. W. Diffusion in paraffin hydrocarbons J. Phys. Chem. 1958, 62 (9) 1102– 1107
- 44Yaws, C. L. Yaws’ Handbook of Physical Properties for Hydrocarbons and Chemicals. http://www.knovel.com/web/portal/browse/display?_EXT_KNOVEL_DISPLAY_bookid=2147 (accessed February 19, 2013) .Google ScholarThere is no corresponding record for this reference.
- 45Tofts, P. S.; Lloyd, D.; Clark, C. A.; Barker, G. J.; Parker, G. J. M.; McConville, P.; Baldock, C.; Pope, J. M. Test liquids for quantitative MRI measurements of self-diffusion coefficient in vivo Magn. Reson. Med. 2000, 43 (3) 368– 374[Crossref], [PubMed], [CAS], Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXitVCgsbc%253D&md5=742855286475e83c3667d5a0bea824baTest liquids for quantitative MRI measurements of self-diffusion coefficient in vivoTofts, P. S.; Lloyd, D.; Clark, C. A.; Barker, G. J.; Parker, G. J. M.; McConville, P.; Baldock, C.; Pope, J. M.Magnetic Resonance in Medicine (2000), 43 (3), 368-374CODEN: MRMEEN; ISSN:0740-3194. (Wiley-Liss, Inc.)A range of liqs. suitable as quality control test objects for measuring the accuracy of clin. MRI diffusion sequences (both apparent diffusion coeff. and tensor) has been identified and characterized. The self-diffusion coeffs. for 15 liqs. (3 cyclic alkanes: cyclohexane to cyclooctane, 9 n-alkanes: n-octane to n-hexadecane, and 3 n-alcs.: ethanol to 1-propanol) were measured at 15-30° using an NMR spectrometer. Values at 22° range from 0.36 to 2.2 10-9 m2s-1. Typical 95% confidence limits are ±2%. Temp. coeffs. are 1.7-3.2%/°C. T1 and T2 values at 1.5 T and proton d. are given. N-tridecane has a diffusion coeff. close to that of normal white matter. The longer n-alkanes may be useful T2 stds. Measurements from a spin-echo MRI sequence agreed to within 2%.
- 46Holler, F.; Callis, J. B. Conformation of the hydrocarbon chains of sodium dodecyl sulfate molecules in micelles: an FTIR study J. Phys. Chem. 1989, 93 (5) 2053– 2058
- 47Lyerla, J. R.; McIntyre, H. M.; Torchia, D. A. A 13C nuclear magnetic resonance study of alkane motion Macromolecules 1974, 7 (1) 11– 14
- 48Jo, S.; Lim, J. B.; Klauda, J. B.; Im, W. CHARMM-GUI membrane builder for mixed bilayers and its application to yeast membranes Biophys. J. 2009, 97 (1) 50– 58[Crossref], [PubMed], [CAS], Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXpsFOmt7g%253D&md5=a6d57a8f42772e4c3674e324ea27ecaaCHARMM-GUI Membrane Builder for mixed bilayers and its application to yeast membranesJo, Sunhwan; Lim, Joseph B.; Klauda, Jeffery B.; Im, WonpilBiophysical Journal (2009), 97 (1), 50-58CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)The CHARMM-GUI Membrane Builder (http://www.charmm-gui.org/input/membrane), an intuitive, straightforward, web-based graphical user interface, was expanded to automate the building process of heterogeneous lipid bilayers, with or without a protein and with support for up to 32 different lipid types. The efficacy of these new features was tested by building and simulating lipid bilayers that resemble yeast membranes, composed of cholesterol, dipalmitoylphosphatidylcholine, dioleoylphosphatidylcholine, palmitoyloleoylphosphatidylethanolamine, palmitoyloleoylphosphatidylamine, and palmitoyloleoylphosphatidylserine. Four membranes with varying concns. of cholesterol and phospholipids were simulated, for a total of 170 ns at 303.15 K. Unsatd. phospholipid chain concn. had the largest influence on membrane properties, such as av. lipid surface area, d. profiles, deuterium order parameters, and cholesterol tilt angle. Simulations with a high concn. of unsatd. chains (73%, membraneunsat) resulted in a significant increase in lipid surface area and a decrease in deuterium order parameters, compared with membranes with a high concn. of satd. chains (60-63%, membranesat). The av. tilt angle of cholesterol with respect to bilayer normal was largest, and the distribution was significantly broader for membraneunsat. Moreover, short-lived cholesterol orientations parallel to the membrane surface existed only for membraneunsat. The membranesat simulations were in a liq.-ordered state, and agree with similar exptl. cholesterol-contg. membranes.
- 49Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water J. Chem. Phys. 1983, 79 (2) 926– 935[Crossref], [CAS], Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 50Joung, I. S.; Cheatham, T. E. Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations J. Phys. Chem. B 2008, 112 (30) 9020– 9041[ACS Full Text
], [CAS], Google Scholar
50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXnvFGqtL4%253D&md5=aa489470ae1c7479bf0911710217bd28Determination of Alkali and Halide Monovalent Ion Parameters for Use in Explicitly Solvated Biomolecular SimulationsJoung, In Suk; Cheatham, Thomas E.Journal of Physical Chemistry B (2008), 112 (30), 9020-9041CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Alkali (Li+, Na+, K+, Rb+, and Cs+) and halide (F-, Cl-, Br-, and I-) ions play an important role in many biol. phenomena, roles that range from stabilization of biomol. structure, to influence on biomol. dynamics, to key physiol. influence on homeostasis and signaling. To properly model ionic interaction and stability in atomistic simulations of biomol. structure, dynamics, folding, catalysis, and function, an accurate model or representation of the monovalent ions is critically necessary. A good model needs to simultaneously reproduce many properties of ions, including their structure, dynamics, solvation, and moreover both the interactions of these ions with each other in the crystal and in soln. and the interactions of ions with other mols. At present, the best force fields for biomols. employ a simple additive, nonpolarizable, and pairwise potential for at. interaction. In this work, the authors describe their efforts to build better models of the monovalent ions within the pairwise Coulombic and 6-12 Lennard-Jones framework, where the models are tuned to balance crystal and soln. properties in Ewald simulations with specific choices of well-known water models. Although it has been clearly demonstrated that truly accurate treatments of ions will require inclusion of nonadditivity and polarizability (particularly with the anions) and ultimately even a quantum mech. treatment, the authors' goal was to simply push the limits of the additive treatments to see if a balanced model could be created. The applied methodol. is general and can be extended to other ions and to polarizable force-field models. The authors' starting point centered on observations from long simulations of biomols. in salt soln. with the AMBER force fields where salt crystals formed well below their soly. limit. The likely cause of the artifact in the AMBER parameters relates to the naive mixing of the Smith and Dang chloride parameters with AMBER-adapted Aqvist cation parameters. To provide a more appropriate balance, the authors reoptimized the parameters of the Lennard-Jones potential for the ions and specific choices of water models. To validate and optimize the parameters, the authors calcd. hydration free energies of the solvated ions and also lattice energies (LE) and lattice consts. (LC) of alkali halide salt crystals. This is the first effort that systematically scans across the Lennard-Jones space (well depth and radius) while balancing ion properties like LE and LC across all pair combinations of the alkali ions and halide ions. The optimization across the entire monovalent series avoids systematic deviations. The ion parameters developed, optimized, and characterized were targeted for use with some of the most commonly used rigid and nonpolarizable water models, specifically TIP3P, TIP4PEW, and SPC/E. In addn. to well reproducing the soln. and crystal properties, the new ion parameters well reproduce binding energies of the ions to water and the radii of the first hydration shells. - 51Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. Numerical Recipes: The Art of Scientific Computing, 3rd ed. ed.; Cambridge University Press: New York, 2007.Google ScholarThere is no corresponding record for this reference.
- 52Pastor, R.; Brooks, B.; Szabo, A. An analysis of the accuracy of Langevin and molecular dynamics algorithms Mol. Phys. 1988, 65 (6) 1409– 1419
- 53Roe, D. R.; Cheatham, T. E. PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data J. Chem. Theory Comput. 2013, 9 (7) 3084– 3095[ACS Full Text
], [CAS], Google Scholar
53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXptFehtr8%253D&md5=6f1bee934f13f180bd7e1feb6b78036dPTRAJ and CPPTRAJ: Software for Processing and Analysis of Molecular Dynamics Trajectory DataRoe, Daniel R.; Cheatham, Thomas E.Journal of Chemical Theory and Computation (2013), 9 (7), 3084-3095CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We describe PTRAJ and its successor CPPTRAJ, two complementary, portable, and freely available computer programs for the anal. and processing of time series of three-dimensional at. positions (i.e., coordinate trajectories) and the data therein derived. Common tools include the ability to manipulate the data to convert among trajectory formats, process groups of trajectories generated with ensemble methods (e.g., replica exchange mol. dynamics), image with periodic boundary conditions, create av. structures, strip subsets of the system, and perform calcns. such as RMS fitting, measuring distances, B-factors, radii of gyration, radial distribution functions, and time correlations, among other actions and analyses. Both the PTRAJ and CPPTRAJ programs and source code are freely available under the GNU General Public License version 3 and are currently distributed within the AmberTools 12 suite of support programs that make up part of the Amber package of computer programs (see http://ambermd.org). This overview describes the general design, features, and history of these two programs, as well as algorithmic improvements and new features available in CPPTRAJ. - 54Kučerka, N.; Liu, Y.; Chu, N.; Petrache, H. I.; Tristram-Nagle, S.; Nagle, J. F. Structure of fully hydrated fluid phase DMPC and DLPC lipid bilayers using X-ray scattering from oriented multilamellar arrays and from unilamellar vesicles Biophys. J. 2005, 88 (4) 2626– 2637
- 55Kučerka, N.; Nieh, M.-P.; Katsaras, J. Fluid phase lipid areas and bilayer thicknesses of commonly used phosphatidylcholines as a function of temperature Biochim. Biophys. Acta 2011, 1808 (11) 2761– 2771
- 56Nagle, J. F. Introductory lecture: Basic quantities in model biomembranes Faraday Discuss. 2013, 161 (0) 11– 29
- 57Braun, A. R.; Sachs, J. N.; Nagle, J. F. Comparing simulations of lipid bilayers to scattering data: The GROMOS 43A1-S3 force field J. Phys. Chem. B 2013, 117 (17) 5065– 5072[ACS Full Text
], [CAS], Google Scholar
57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXltl2gs74%253D&md5=7d26d77825a6b983455aed5dac8321a9Comparing Simulations of Lipid Bilayers to Scattering Data: The GROMOS 43A1-S3 Force FieldBraun, Anthony R.; Sachs, Jonathan N.; Nagle, John F.Journal of Physical Chemistry B (2013), 117 (17), 5065-5072CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Simulations of DOPC at T = 303 K were performed using the united atom force field 43A1-S3 at six fixed projected areas, AP = 62, 64, 66, 68, 70, and 72 Å2, as well as a tensionless simulation that produced an av. ANPT = 65.8 Å2. After a small undulation correction for the system size consisting of 288 lipids, results were compared to exptl. data. The best, and excellent, fit to neutron scattering data occurs at an interpolated AN = 66.6 Å2 and the best, but not as good, fit to the more extensive x-ray scattering data occurs at AX = 68.7 Å2. The distance ΔDB-H between the Gibbs dividing surface for water and the peak in the electron d. profile agrees with scattering expts. The calcd. area compressibility KA = 277±10 mN/m is in excellent agreement with the micromech. expt. The vol. per lipid VL is smaller than vol. expts. which suggests a workaround that raises all the areas by about 1.5%. Although AX ≠ AN ≠ ANPT, this force field obtains acceptable agreement with expt. for AL = 67.5 Å2 (68.5 Å2 in the workaround), which we suggest is a better DOPC result from 43A1-S3 simulations than its value from the tensionless NPT simulation. However, nonsimulation modeling obtains better simultaneous fits to both kinds of scattering data, which suggests that the force fields can still be improved. - 58Rawicz, W.; Olbrich, K. C.; McIntosh, T.; Needham, D.; Evans, E. Effect of chain length and unsaturation on elasticity of lipid bilayers Biophys. J. 2000, 79 (1) 328– 339[Crossref], [PubMed], [CAS], Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1Kksrs%253D&md5=5d0a8c5e727c7620b846aebd6cc97cf5Effect of chain length and unsaturation on elasticity of lipid bilayersRawicz, W.; Olbrich, K. C.; McIntosh, T.; Needham, D.; Evans, E.Biophysical Journal (2000), 79 (1), 328-339CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Micropipette pressurization of giant bilayer vesicles was used to measure both elastic bending kc and area stretch KA moduli of fluid-phase phosphatidylcholine (PC) membranes. Twelve diacyl PCs were chosen: eight with two 18 carbon chains and degrees of unsatn. from one double bond (C18:1/0, C18:0/1) to six double bonds per lipid (diC18:3), two with short satd. carbon chains (diC13:0, diC14:0), and two with long unsatd. carbon chains (diC20:4, diC22:1). Bending moduli were derived from measurements of apparent expansion in vesicle surface area under very low tensions (0.001-0.5 mN/m), which is dominated by smoothing of thermal bending undulations. Area stretch moduli were obtained from measurements of vesicle surface expansion under high tensions (>0.5 mN/m), which involve an increase in area per mol. and a small, but important, contribution from smoothing of residual thermal undulations. The direct stretch moduli varied little (< ± 10%) with either chain unsatn. or length about a mean of 243 mN/m. On the other hand, the bending moduli of satd./mono-unsatd. chain PCs increased progressively with chain length from 0.56 × 10-19 J for diC13:0 to 1.2 × 10-19 J for diC22:1. However, quite unexpectedly for longer chains, the bending moduli dropped precipitously to ∼0.4 × 10-19 J when two or more cis double bonds were present in a chain (C18:0/2, diC18:2, diC18:3, diC20:4). Given nearly const. area stretch moduli, the variations in bending rigidity with chain length and poly-unsatn. implied significant variations in thickness. To test this hypothesis, peak-to-peak headgroup thicknesses hpp of bilayers were obtained from x-ray diffraction of multibilayer arrays at controlled relative humidities. For satd./mono-unsatd. chain bilayers, the distances hpp increased smoothly from diC13:0 to diC22:1 as expected. Moreover, the distances and elastic properties correlated well with a polymer brush model of the bilayer that specifies that the elastic ratio (kc/KA)1/2 = (hpp - ho)/24, where ho ≈ 1 nm accounts for sepn. of the headgroup peaks from the deformable hydrocarbon region. However, the elastic ratios and thicknesses for diC18:2, diC18:3, and diC20:4 fell into a distinct group below the correlation, which showed that poly-cis unsatd. chain bilayers are thinner and more flexible than satd./mono-unsatd. chain bilayers.
- 59Petrache, H. I.; Tristram-Nagle, S.; Nagle, J. F. Fluid phase structure of EPC and DMPC bilayers Chem. Phys. Lipids 1998, 95 (1) 83– 94
- 60Klauda, J. B.; Kučerka, N.; Brooks, B. R.; Pastor, R. W.; Nagle, J. F. Simulation-based methods for interpreting X-ray data from lipid bilayers Biophys. J. 2006, 90 (8) 2796– 2807[Crossref], [PubMed], [CAS], Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjsFeqtbw%253D&md5=727cd5ebed91aa22a6c6d1bff8000dd6Simulation-based methods for interpreting X-ray data from lipid bilayersKlauda, Jeffery B.; Kucerka, Norbert; Brooks, Bernard R.; Pastor, Richard W.; Nagle, John F.Biophysical Journal (2006), 90 (8), 2796-2807CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)The fully hydrated liq. cryst. phase of the dimyristoylphosphatidylcholine lipid bilayer at 30° was simulated using mol. dynamics with the CHARMM potential for five surface areas per lipid (A) in the range 55-65 Å2 that brackets the previously detd. exptl. area 60.6 Å2. The results of these simulations are used to develop a new hybrid zero-baseline structural model, denoted H2, for the electron d. profile, p(z), for the purpose of interpreting x-ray diffraction data. H2 and also the older hybrid baseline model were tested by fitting to partial information from the simulation and various constraints, both of which correspond to those available exptl. The A, ρ(z), and F(q) obtained from the models agree with those calcd. directly from simulation at each of the five areas, thereby validating this use of the models. The new H2 was then applied to exptl. dimyristoylphosphatidylcholine data; it yields A = 60.6 ± 0.5 Å2, in agreement with the earlier est. obtained using the hybrid baseline model. The electron d. profiles also compare well, despite considerable differences in the functional forms of the two models. Overall, the simulated ρ(z) at A = 60.7 Å2 agrees well with expt., demonstrating the accuracy of the CHARMM lipid force field; small discrepancies indicate targets for improvements. Lastly, a simulation-based model-free approach for obtaining A is proposed. It is based on interpolating the area that minimizes the difference between the exptl. F(q) and simulated F(q) evaluated for a range of surface areas. This approach is independent of structural models and could be used to det. structural properties of bilayers with different lipids, cholesterol, and peptides.
- 61Kučerka, N.; Tristram-Nagle, S.; Nagle, J. F. Closer look at structure of fully hydrated fluid phase DPPC bilayers Biophys. J. 2006, 90 (11) L83– L85
- 62Kučerka, N.; Nagle, J. F.; Sachs, J. N.; Feller, S. E.; Pencer, J.; Jackson, A.; Katsaras, J. Lipid bilayer structure determined by the simultaneous analysis of neutron and X-ray scattering data Biophys. J. 2008, 95 (5) 2356– 2367
- 63Evans, E.; Rawicz, W.; Smith, B. A. Concluding remarks back to the future: mechanics and thermodynamics of lipid biomembranes Faraday Discuss. 2013, 161 (0) 591– 611
- 64Evans, E.Personal Communication - DOPC isothermal compressibility modulus from X-ray data at 293 K; 2014.Google ScholarThere is no corresponding record for this reference.
- 65Tristram-Nagle, S.; Petrache, H. I.; Nagle, J. F. Structure and interactions of fully hydrated dioleoylphosphatidylcholine bilayers Biophys. J. 1998, 75 (2) 917– 925
- 66Pan, J.; Tristram-Nagle, S.; Kučerka, N.; Nagle, J. F. Temperature dependence of structure, bending rigidity, and bilayer interactions of dioleoylphosphatidylcholine bilayers Biophys. J. 2008, 94 (1) 117– 124
- 67Liu, Y.; Nagle, J. F. Diffuse scattering provides material parameters and electron density profiles of biomembranes Phys. Rev. E 2004, 69 (4) 040901
- 68Kučerka, N.; Tristram-Nagle, S.; Nagle, J. F. Structure of fully hydrated fluid phase lipid bilayers with monounsaturated chains J. Membr. Biol. 2006, 208 (3) 193– 202[Crossref], [CAS], Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjvF2lsbY%253D&md5=734e76569cbe3cf00f2a51d2a9543b7eStructure of Fully Hydrated Fluid Phase Lipid Bilayers with Monounsaturated ChainsKucerka, Norbert; Tristram-Nagle, Stephanie; Nagle, John F.Journal of Membrane Biology (2006), 208 (3), 193-202CODEN: JMBBBO; ISSN:0022-2631. (Springer)Quant. structures are obtained at 30°C for the fully hydrated fluid phases of palmitoyloleoylphosphatidylcholine (POPC), with a double bond on the sn-2 hydrocarbon chain, and for dierucoylphosphatidylcholine (di22:1PC), with a double bond on each hydrocarbon chain. The form factors F(qz) for both lipids are obtained using a combination of 3 methods: (1) Volumetric measurements provide F(0), (2) x-ray scattering from extruded unilamellar vesicles provides |F(qz)| for low qz, (3) Diffuse x-ray scattering from oriented stacks of bilayers provides |F(qz)| for high qz. Also, data using method (2) are added to the authors' recent data for dioleoylphosphatidylcholine (DOPC) using methods (1) and (3); the new DOPC data agree very well with the recent data and with (4) the authors' older data obtained using a liq. crystallog. x-ray method. The authors used hybrid electron d. models to obtain structural results from these form factors. The result for area per lipid (A) for DOPC 72.4 ± 0.5 Å2 agrees well with the authors' earlier publications, and the authors find A = 69.3 ± 0.5 Å2 for di22:1PC and A = 68.3 ± 1.5 Å2 for POPC. The authors obtain the values for 5 different av. thicknesses: hydrophobic, steric, head-head, phosphate-phosphate and Luzzati. Comparison of the results for these 3 lipids and for the authors' recent dimyristoylphosphatidylcholine (DMPC) detn. provides quant. measures of the effect of unsatn. on bilayer structure. The authors' results suggest that lipids with one monounsatd. chain have quant. bilayer structures closer to lipids with 2 monounsatd. chains than to lipids with 2 completely satd. chains.
- 69Binder, H.; Gawrisch, K. Effect of unsaturated lipid chains on dimensions, molecular order and hydration of membranes J. Phys. Chem. B 2001, 105 (49) 12378– 12390
- 70Rand, R. P.; Parsegian, V. A. Hydration forces between phospholipid bilayers Biochim. Biophys. Acta 1989, 988 (3) 351– 376
- 71Rappolt, M.; Hickel, A.; Bringezu, F.; Lohner, K. Mechanism of the lamellar/inverse hexagonal phase transition examined by high resolution X-ray diffraction Biophys. J. 2003, 84 (5) 3111– 3122[Crossref], [PubMed], [CAS], Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjvFGjur4%253D&md5=6a145c30169a2a14b0e664a625fbcf00Mechanism of the lamellar/inverse hexagonal phase transition examined by high resolution X-ray diffractionRappolt, Michael; Hickel, Andrea; Bringezu, Frank; Lohner, KarlBiophysical Journal (2003), 84 (5), 3111-3122CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)For the first time the electron d. of the lamellar liq. cryst. as well as of the inverted hexagonal phase could be retrieved at the transition temp. A reliable decompn. of the d-spacings into hydrophobic and hydrophilic structure elements could be performed owing to the presence of a sufficient no. of reflections. While the hydrocarbon chain length, dC, in the lamellar phase with a value of 14.5 Å lies within the extreme limits of the estd. chain length of the inverse hexagonal phase 10 Å < dC < 16 Å, the changes in the hydrophilic region vary strongly. During the lamellar-to-inverse hexagonal phase transition the area per lipid mol. reduces by ∼25%, and the no. of water mols. per lipid increases from 14 to 18. On the basis of the anal. of the structural components of each phase, the interface between the coexisting mesophases between 66 and 84° has been examd. in detail, and a model for the formation of the first rods in the matrix of the lamellar phospholipid stack is discussed. Judging from the structural relations between the inverse hexagonal and the lamellar phase, we suggest a cooperative chain reaction of rod formation at the transition midpoint, which is mainly driven by minimizing the interstitial region.
- 72Nagle, J. F.; Tristram-Nagle, S. Lipid bilayer structure Curr. Opin. Struct. Biol. 2000, 10 (4) 474– 480[Crossref], [PubMed], [CAS], Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXlvFOmt7Y%253D&md5=9cf3b9383d43db17029625724053da1eLipid bilayer structureNagle, John F.; Tristram-Nagle, StephanieCurrent Opinion in Structural Biology (2000), 10 (4), 474-480CODEN: COSBEF; ISSN:0959-440X. (Elsevier Science Ltd.)A review with 51 refs. Fluctuations, inherent in flexible and biol. relevant lipid bilayers, make quant. structure detn. challenging. Shortcomings in older methods of structure detn. have been realized and new methodologies have been introduced that take fluctuations into account. The large uncertainty in literature values of lipid bilayer structural parameters is being reduced.
- 73Anézo, C.; de Vries, A. H.; Höltje, H.-D.; Tieleman, D. P.; Marrink, S.-J. Methodological issues in lipid bilayer simulations J. Phys. Chem. B 2003, 107 (35) 9424– 9433
- 74Poger, D.; Mark, A. E. On the validation of molecular dynamics simulations of saturated and cis-monounsaturated phosphatidylcholine lipid bilayers: A comparison with experiment J. Chem. Theory Comput. 2009, 6 (1) 325– 336
- 75Kučerka, N.; Katsaras, J.; Nagle, J. Comparing membrane simulations to scattering experiments: Introducing the SIMtoEXP software J. Membr. Biol. 2010, 235 (1) 43– 50[Crossref], [PubMed], [CAS], Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXls1Wkurg%253D&md5=52e94861b62db6b538f7dc502b680818Comparing Membrane Simulations to Scattering Experiments: Introducing the SIMtoEXP SoftwareKucerka, Norbert; Katsaras, John; Nagle, John F.Journal of Membrane Biology (2010), 235 (1), 43-50CODEN: JMBBBO; ISSN:0022-2631. (Springer)SIMtoEXP is a software package designed to facilitate the comparison of biomembrane simulations with exptl. x-ray and neutron scattering data. It has the following features: (1) Accepts no. d. profiles from simulations in a std. but flexible format. (2) Calcs. the electron d. ε(z) and neutron scattering length d. ν(z) profiles along the z direction (i.e., normal to the membrane) and their resp. Fourier transforms (i.e., Fε [qz] and Fν[qz]). The resultant 4 functions are then displayed graphically. (3) Accepts exptl. Fε(qz) and Fν(qz) data for graphical comparison with simulations. (4) Allows for lipids and other large mols. to be parsed into component groups by the user and calcs. the component vols. following Petrache et al. The software then calcs. and displays the contributions of each component group as vol. probability profiles, ρ(z), as well as the contributions of each component to ε(z) and ν(z).
- 76Shirts, M. R. Simple quantitative tests to validate sampling from thermodynamic ensembles J. Chem. Theory Comput. 2012, 9 (2) 909– 926
- 77Seelig, J.; Waespe-Sarcevic, N. Molecular order in cis and trans unsaturated phospholipid bilayers Biochemistry 1978, 17 (16) 3310– 3315
- 78Perly, B.; Smith, I. C. P.; Jarrell, H. C. Acyl chain dynamics of phosphatidylethanolamines containing oleic acid and dihydrosterculic acid: deuteron NMR relaxation studies Biochemistry 1985, 24 (17) 4659– 4665
- 79Lafleur, M.; Bloom, M.; Eikenberry, E. F.; Gruner, S. M.; Han, Y.; Cullis, P. R. Correlation between lipid plane curvature and lipid chain order Biophys. J. 1996, 70 (6) 2747– 2757[Crossref], [PubMed], [CAS], Google Scholar79https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVCrtb4%253D&md5=ff34b0b6df6c340bdc716572baf34935Correlation between lipid plane curvature and lipid chain orderLafleur, Michel; Bloom, Myer; Eikenberry, Eric F.; Gruner, Sol M.; Han, Yuqi; Cullis, Pieter R.Biophysical Journal (1996), 70 (6), 2747-2757CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)The 1-palmitoyl-2-oleoyl-phosphatidylethanolamine: 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPE:POPC) system has been investigated by measuring, in the inverted hexagonal (HII) phase, the intercylinder spacings (using x-ray diffraction) and orientational order of the acyl chains (using 2H NMR). The presence of 20% dodecane leads to the formation of a HII phase for the compn. range from 0 to 39 mol% of POPC in POPE, as ascertained by x-ray diffraction and 2H NMR. The addn. of the alkane induces a small decrease in chain order, consistent with less stretched chains. An increase in temp. or in POPE proportion leads to a redn. in the intercylinder spacing, primarily due to a decrease in the water core radius. A temp. increase also leads to a redn. in the orientational order of the lipid acyl chains, whereas in POPE proportion has little effect on chain order. A correlation is proposed to relate the radius of curvature of the cylinders in the inverted hexagonal phase to the chain order of the lipids adopting the HII phase. A simple geometrical model is proposed, taking into account the area occupied by the polar headgroup at the interface and the orientational order of the acyl chains reflecting the contribution of the apolar core. From these parameters, intercylinder spacings are calcd. that agree well with the values detd. exptl. by x-ray diffraction, for the variations of both temp. and POPE:POPC proportion. This model suggests that temp. increases the curvature of lipid layers, mainly by increasing the area subtended by the hydrophobic core through chain conformation disorder, whereas POPC content affects primarily the headgroup interface contribution. The frustration of lipid layer curvature is also shown to be reflected in the acyl chain order measured in the Lα phase, in the absence of dodecane; for a given temp., increased order is obsd. when the curling tendencies of the lipid plane are more pronounced.
- 80Warschawski, D.; Devaux, P. Order parameters of unsaturated phospholipids in membranes and the effect of cholesterol: a 1H-13C solid-state NMR study at natural abundance Eur. Biophys. J. 2005, 34 (8) 987– 96[Crossref], [PubMed], [CAS], Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SksbvJ&md5=b5ac0c99dd2dad942a2c48c20e8516c5Order parameters of unsaturated phospholipids in membranes and the effect of cholesterol: a 1H-13C solid-state NMR study at natural abundanceWarschawski, Dror E.; Devaux, Philippe F.European Biophysics Journal (2005), 34 (8), 987-996CODEN: EBJOE8; ISSN:0175-7571. (Springer)Most biol. phospholipids contain at least one unsatd. alkyl chain. However, few order parameters of unsatd. lipids have been detd. because of the difficulty assocd. with isotopic labeling of a double bond. Dipolar recoupling on axis with scaling and shape preservation (DROSS) is a solid-state NMR technique optimized for measuring 1H-13C dipolar couplings and order parameters in lipid membranes in the fluid phase. It has been used to det. the order profile of 1,2-dimyristoyl-sn-glycero-3-phosphocholine hydrated membranes. Here, we show an application for the measurement of local order parameters in multilamellar vesicles contg. unsatd. lipids. Taking advantage of the very good 13C chem. shift dispersion, one can easily follow the segmental order along the acyl chains and, particularly, around the double bonds where we have been able to det. the previously misassigned order parameters of each acyl chain of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC). We have followed the variation of such order profiles with temp., unsatn. content and cholesterol addn. We have found that the phase formed by DOPC with 30% cholesterol is analogous to the liq.-ordered (lo) phase. Because these expts. do not require isotopic enrichment, this technique can, in principle, be applied to natural lipids and biomembranes.
- 81Petrache, H. I.; Dodd, S. W.; Brown, M. F. Area per lipid and acyl length distributions in fluid phosphatidylcholines determined by (2)H NMR spectroscopy Biophys. J. 2000, 79 (6) 3172– 92[Crossref], [PubMed], [CAS], Google Scholar81https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXitFCn&md5=26c87135c39ab21617ea1987653b083fArea per lipid and acyl length distributions in fluid phosphatidylcholines determined by 2H NMR spectroscopyPetrache, Horia I.; Dodd, Steven W.; Brown, Michael F.Biophysical Journal (2000), 79 (6), 3172-3192CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Deuterium (2H) NMR spectroscopy provides detailed information regarding the structural fluctuations of lipid bilayers, including both the equil. properties and dynamics. Exptl. 2H NMR measurements for the homologous series of 1,2-diacyl-sn-glycero-3-phosphocholines with perdeuterated satd. chains (from C12:0 to C18:0) have been performed on randomly oriented, fully hydrated multilamellar samples. For each lipid, the C-D bond order parameters have been calcd. from de-Paked 2H NMR spectra as a function of temp. The exptl. order parameters were analyzed using a mean-torque potential model for the acyl chain segment distributions, and comparison was made with the conventional diamond lattice approach. Statistical mech. principles were used to relate the measured order parameters to the lipid bilayer structural parameters: the hydrocarbon thickness and the mean interfacial area per lipid. At fixed temp., the area decreases with increasing acyl length, indicating increased van der Waals attraction for longer lipid chains. However, the main effect of increasing the acyl chain length is on the hydrocarbon thickness rather than on the area per lipid. Expansion coeffs. of the structural parameters are reported and interpreted using an empirical free energy function that describes the force balance in fluid bilayers. At the same abs. temp., the phosphatidylcholine (PC) series exhibits a universal chain packing profile that differs from that of phosphatidylethanolamines (PE). Hence, the lateral packing of phospholipids is more sensitive to the headgroup methylation than to the acyl chain length. A fit to the area per lipid for the PC series using the empirical free energy function shows that the PE area represents a limiting value for the packing of fluid acyl chains.
- 82Douliez, J. P.; Léonard, A.; Dufourc, E. J. Restatement of order parameters in biomembranes: calculation of C-C bond order parameters from C-D quadrupolar splittings Biophys. J. 1995, 68 (5) 1727– 1739[Crossref], [PubMed], [CAS], Google Scholar82https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXlt1KgsrY%253D&md5=cb2fa7e7b26000ab88e390ee0bd25090Restatement of order parameters in biomembranes: calculation of C-C bond order parameters from C-D quadrupolar splittingsDouliez, Jean-Paul; Leonard, Alain; Dufourc, Erick J.Biophysical Journal (1995), 68 (5), 1727-39CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)An expression for the C-C bond order parameter, SCC, of membrane hydrocarbon chains has been derived from the obsd. C-D bond order parameters. It allows calcn. of the probability of each of the C-C bond rotamers and, consequently, the no. of gauche defects per chain as well as their projected av. length onto the bilayer normal, thus affording the calcn. of accurate hydrophobic bilayer thicknesses. The effect of the temp. has been studied on dilauroyl-, dimyristoyl-, and dipalmitoylphosphatidylcholine (DLPC, DMPC, DPPC) membranes, as has the effect of cholesterol on DMPC. The salient results are as follows: (1) an odd-even effect is obsd. for the SCC vs. carbon position, k, whose amplitude increases with temp.; (2) calcn. of SCC, from nonequivalent deuterons on the sn-2 chain of lipids, S2CC, leads to neg. values, indicating the tendency for the C1-C2 bond to be oriented parallel to the bilayer surface; this bond becomes more parallel to the surface as the temp. increases or when cholesterol is added; (3) calcn. on the sn-2 chain length can be performed from C1 and Cn, where n is the no. of carbon atoms in the chain, and leads to 10.4, 12.2, and 13.8 Å for DLPC, DMPC, and DPPC close to the transition temp. TC, of each of the systems and to 9.4, 10.9, and 12.6 for T-TC = 30-40°, resp.; (4) sepn. of intra- and intermol. motions allows quantitation of the no. of gauche defects per chain, which is equal to 1.9, 2.7, and 3.5 for DLPC, DMPC, and DPPC near Tc and to 2.7, 3.5, and 4.4 at T - TC = 30-40°, resp. Finally, the validity of the model is discussed and compared with previously published models.
- 83Aussenac, F.; Laguerre, M.; Schmitter, J.-M.; Dufourc, E. J. Detailed structure and dynamics of bicelle phospholipids using selectively deuterated and perdeuterated labels. 2H NMR and molecular mechanics study Langmuir 2003, 19 (25) 10468– 10479
- 84Shaikh, S. R.; Brzustowicz, M. R.; Gustafson, N.; Stillwell, W.; Wassall, S. R. Monounsaturated PE does not phase-separate from the lipid raft molecules sphingomyelin and cholesterol: Role for polyunsaturation? Biochemistry 2002, 41 (34) 10593– 10602
- 85Hitchcock, P. B.; Mason, R.; Thomas, K. M.; Shipley, G. G. Structural chemistry of 1,2 dilauroyl-DL-phosphatidylethanolamine: Molecular conformation and intermolecular packing of phospholipids Proc. Natl. Acad. Sci. U.S.A. 1974, 71 (8) 3036– 3040
- 86Seelig, A.; Seelig, J. Bilayers of dipalmitoyl-3-sn-phosphatidylcholine: Conformational differences between the fatty acyl chains Biochim. Biophys. Acta 1975, 406 (1) 1– 5
- 87Jämbeck, J. P. M.; Lyubartsev, A. P. An extension and further validation of an all-atomistic force field for biological membranes J. Chem. Theory Comput. 2012, 8 (8) 2938– 2948
- 88Senak, L.; Davies, M. A.; Mendelsohn, R. A quantitative IR study of hydrocarbon chain conformation in alkanes and phospholipids: CH2 wagging modes in disordered bilayer and HII phases J. Phys. Chem. 1991, 95 (6) 2565– 2571[ACS Full Text
], [CAS], Google Scholar
88https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhvFemsLw%253D&md5=86298b46c827e9371a74b49f9dd28aefA quantitative IR study of hydrocarbon chain conformation in alkanes and phospholipids: CH2 wagging modes in disordered bilayer and HII phasesSenak, Laurence; Davies, Mark A.; Mendelsohn, RichardJournal of Physical Chemistry (1991), 95 (6), 2565-71CODEN: JPCHAX; ISSN:0022-3654.A series of CH2 wagging modes in the 1320-1370-cm-1 region of alkane and phospholipid IR spectra characteristic of nonplanar conformers were used for quant. evaluation of conformational states in disordered (phospholipid Lα and HII and alkane liq.) phases. A quant. comparison of DPPC and 1,2-dipalmitoylphosphatidylethanolamine (DPPE) shows the former to have 0.4 double gauche (gg), 0.5 end gauche (eg), and ∼1.0 (kink + gtg) conformers per chain just above the gel0liq. crystal phase transition, while the more highly ordered DPPE Lα phase shows ∼0.2 gg, 0.1 eg, and 1.0 (kink + gtg) conformers. In all systems studied, the no. of allowed gg and eg forms that occur is reduced substantially. From those achieved in isotropic liq. alkanes. A study of the Lα → HII interconversion in 2 unsatd. (1-palmitoyl-2-oleoyl- and 1,2-dielaidoylphosphatidylethanolamine) shows a substantial increase in both gg and eg conformers near temp.3s leading to the inverted micellar state in each instance, with smaller percentage increases in (kink + gtg) states. Connections between these quant. observations with those of 2H NMR and other spectroscopies are presented. Finally, a temp.4 dependence inconsistent with the prediction of the rotational isomeric state model is noted for the 1368-cm-1 band characteristic of (kink + gtg) conformers in alkanes. - 89Moss, G. P. Basic terminology of stereochemistry Pure Appl. Chem. 1996, 68 (12) 2193– 2222
- 90Cates, D. A.; Strauss, H. L.; Snyder, R. G. Vibrational modes of liquid n-alkanes: Simulated isotropic raman spectra and band progressions for C5H12-C20H42 and C16D34 J. Phys. Chem. 1994, 98 (16) 4482– 4488
- 91Snyder, R. G.; Strauss, H. L.; Elliger, C. A. C-H stretching modes and the structure of n-alkyl chains. 1. Long, disordered chains J. Phys. Chem. 1982, 86, 5145– 5150
- 92Mendelsohn, R.; Senak, L. Quantitative determination of conformational disorder in biological membranes by FTIR spectroscopy. In Biomolecular spectroscopy; Clark, R. J. H.; Hester, R. E., Eds.; Wiley: New York, 1993; pp 339– 380.Google ScholarThere is no corresponding record for this reference.
- 93Casal, H. L.; McElhaney, R. N. Quantitative determination of hydrocarbon chain conformational order in bilayers of saturated phosphatidylcholines of various chain lengths by Fourier transform infrared spectroscopy Biochemistry 1990, 29 (23) 5423– 5427
- 94Tuchtenhagen, J.; Ziegler, W.; Blume, A. Acyl chain conformational ordering in liquid-crystalline bilayers: comparative FT-IR and 2H-NMR studies of phospholipids differing in headgroup structure and chain length Eur. Biophys. J. 1994, 23 (5) 323– 335[Crossref], [CAS], Google Scholar94https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXltFClurg%253D&md5=44dcd68c3e24ac87b38aacf1dc9a1338Acyl chain conformational ordering in liquid-crystalline bilayers: comparative FT-IR and 2H-NMR studies of phospholipids differing in headgroup structure and chain lengthTuchtenhagen, Juergen; Ziegler, Wolfgang; Blume, AlfredEuropean Biophysics Journal (1994), 23 (5), 323-35CODEN: EBJOE8; ISSN:0175-7571.FT-IR spectroscopy has been used to evaluate the acyl chain conformational ordering of DMPC, DMPE, DMPA (pH 6 and 12), DMPG (pH 1 and 7), and DPPC, DPPE, DPPA (pH 6). The frequencies of the sym. and antisym. methylene stretching vibrations were detd. as a function of temp. In the liq.-cryst. phase the frequencies show a qual. dependence on the amt. of chain disorder. Quant. data for trans-gauche isomerization were obtained from the integral intensities of the conformation sensitive methylene wagging absorptions at ∼1368 cm-1 (gtg' and gtg sequences), 1356 cm-1 (double gauche) and 1342 cm-1 (end gauche). The integral band intensities were converted to the no. of gauche conformers per acyl chain using the calibration factors published by L. Senak et al. (1991). At 69° the highest no. of gauche conformers excluding contributions from single gauche conformers and jogs (g'tttg) were found for PCs (DMPC: 2.6; DPPC: 2.4), followed by DMPG- (2.0), phosphatidylethanolamines (DMPE: 1.4; DPPE: 2.0), protonated DMPG (1.5), and phosphatidic acids (DPPA-: 1.7; DMPA-: 1.4, DMPA2-: 1.7).
- 95Mendelsohn, R.; Davies, M. A.; Brauner, J. W.; Schuster, H. F.; Dluhy, R. A. Quantitative determination of conformational disorder in the acyl chains of phospholipid bilayers by infrared spectroscopy Biochemistry 1989, 28 (22) 8934– 8939
- 96Kučerka, N.; Gallová, J.; Uhríková, D.; Balgavý, P.; Bulacu, M.; Marrink, S.-J.; Katsaras, J. Areas of Monounsaturated Diacylphosphatidylcholines Biophys. J. 2009, 97 (7) 1926– 1932
- 97Ratto, T. V.; Longo, M. L. Obstructed diffusion in phase-separated supported lipid bilayers: A combined atomic force microscopy and fluorescence recovery after photobleaching approach Biophys. J. 2002, 83 (6) 3380– 3392
- 98Almeida, P. F. F.; Vaz, W. L. C.; Thompson, T. E. Lateral diffusion in the liquid phases of dimyristoylphosphatidylcholine/cholesterol lipid bilayers: a free volume analysis Biochemistry 1992, 31 (29) 6739– 6747[ACS Full Text
], [CAS], Google Scholar
98https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38Xks1Gisbc%253D&md5=1478036ef44a66d02c70da8e2dd0ff8cLateral diffusion in the liquid phases of dimyristoylphosphatidylcholine/cholesterol lipid bilayers: a free volume analysisAlmeida, Paulo F. F.; Vaz, Winchil L. C.; Thompson, T. E.Biochemistry (1992), 31 (29), 6739-47CODEN: BICHAW; ISSN:0006-2960.Fluorescence recovery after photobleaching is used to perform an extensive study of the lateral diffusion of a phospholipid probe in the binary mixt. dimyristoylphosphatidylcholine/cholesterol, above the melting temp. of the phospholipid. In the regions of the phase diagram where a single liq. phase exists, diffusion can be quant. described by free vol. theory, using a modified Macedo-Litovitz hybrid equation. In the liq.-liq. immiscibility region, the temp. dependence of the diffusion coeff. is in excellent agreement with current theories of generalized diffusivities in composite two-phase media. A consistent interpretation of the diffusion data can be provided based essentially on the idea that the primary effect of cholesterol addn. to the bilayer is to occupy free vol. On this basis, a general interpretation of the phase behavior of this mixt. is also proposed. - 99Orädd, G.; Lindblom, G.; Westerman, P. W. Lateral diffusion of cholesterol and dimyristoylphosphatidylcholine in a lipid bilayer measured by pulsed field gradient NMR spectroscopy Biophys. J. 2002, 83 (5) 2702– 2704
- 100Filippov, A.; Orädd, G.; Lindblom, G. Influence of cholesterol and water content on phospholipid lateral diffusion in bilayers Langmuir 2003, 19 (16) 6397– 6400
- 101Vaz, W. L. C.; Clegg, R. M.; Hallmann, D. Translational diffusion of lipids in liquid crystalline phase phosphatidylcholine multibilayers. A comparison of experiment with theory Biochemistry 1985, 24 (3) 781– 786[ACS Full Text
], [CAS], Google Scholar
101https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXptFKqsA%253D%253D&md5=5c01282a25df5150530f13a60cbc9dddTranslational diffusion of lipids in liquid crystalline phase phosphatidylcholine multibilayers. A comparison of experiment with theoryVaz, Winchil L. C.; Clegg, Robert M.; Hallmann, DieterBiochemistry (1985), 24 (3), 781-6CODEN: BICHAW; ISSN:0006-2960.A systematic study of the translational diffusion of the phospholipid deriv., N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine (NBD-PE), was undertaken in liq. cryst. phase phosphatidylcholine (PC) bilayers by using the fluorescence recovery after photobleaching technique. The exptl. results were then compared with the predictions of theor. models for diffusion in membranes. For NBD-PE, the dependence of the translational diffusion coeff. (Dt) upon the acyl chain length of the diffusant was not that predicted by continuum fluid hydrodynamic models for diffusion in membranes. Plots of Dt vs. 1/T (Arrhenius plots) were nonlinear in PC bilayers where the acyl chain compn. of the NBD-PE was matched with that of the host bilayer lipid. This suggested that a free vol. model may be appropriate for the description of lipid diffusion in lipid bilayers. In bilayers of PCs with satd. acyl chains at the same reduced temp., the magnitude of Dt followed the order: distearoyl-PC > dipalmitoyl-PC (DPPC) > dimyristoyl-PC (DMPC) > dilauroyl-PC (DLPC). This was the inverse of what might be expected from the hydrodynamic model, but was in agreement with the free vol. in these bilayers. A free vol. model that takes into account the frictional drag forces acting upon the diffusing NBD-PE at the membrane-water interface and also at the bilayer midplane adequately describes the diffusion results for NBD-PE mols. in DLPC, DMPC, DPPC, and 1-palmitoyl-2-oleoyl-PC bilayers in the liq.-cryst. phase. - 102Scheidt, H. A.; Huster, D.; Gawrisch, K. Diffusion of cholesterol and its precursors in lipid membranes studied by 1H pulsed field gradient magic angle spinning NMR Biophys. J. 2005, 89 (4) 2504– 2512
- 103Kuśba, J.; Li, L.; Gryczynski, I.; Piszczek, G.; Johnson, M.; Lakowicz, J. R. Lateral diffusion coefficients in membranes measured by resonance energy transfer and a new algorithm for diffusion in two dimensions Biophys. J. 2002, 82 (3) 1358– 1372
- 104Jin, A. J.; Edidin, M.; Nossal, R.; Gershfeld, N. L. A singular state of membrane lipids at cell growth temperatures Biochemistry 1999, 38 (40) 13275– 13278[ACS Full Text
], [CAS], Google Scholar
104https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXlvVSisrs%253D&md5=ae149f4a53ff9c3f3e906f6763212046A singular state of membrane lipids at cell growth temperaturesJin, A. J.; Edidin, M.; Nossal, R.; Gershfeld, N. L.Biochemistry (1999), 38 (40), 13275-13278CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Cells adjust their membrane lipid compn. when they adapt to grow at different temps. The consequences of this adjustment for membrane properties and functions are not well understood. Our report shows that the temp. dependence of the diffusion of a probe mol. in multilayers formed from total lipid exts. of E. coli has an anomalous max. at a temp. corresponding to the growth temp. of each bacterial prepn. (25, 29, and 32 °C). This increase in the lateral diffusion coeff., D, is characteristic of membrane lipids in a crit. state, for which large fluctuations of mol. area in the plane of the bilayer are expected. Therefore, changes in lipid compn. may be due to a requirement that cells maintain their membranes in a state where mol. interactions and reaction rates are readily modulated by small, local perturbations of membrane organization. - 105Poger, D.; Mark, A. E. Lipid bilayers: The effect of force field on ordering and dynamics J. Chem. Theory Comput. 2012, 8 (11) 4807– 4817[ACS Full Text
], [CAS], Google Scholar
105https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVOhs7vI&md5=2870d3c59c4b8baa96dae55ca46b3a75Lipid Bilayers: The Effect of Force Field on Ordering and DynamicsPoger, David; Mark, Alan E.Journal of Chemical Theory and Computation (2012), 8 (11), 4807-4817CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The sensitivity of the structure and dynamics of a fully hydrated pure bilayer of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) in mol. dynamics simulations to changes in force-field and simulation parameters has been assessed. Three related force fields (the GROMOS 54A7 force field, a GROMOS 53A6-derived parameter set and a variant of the Berger parameters) in combination with either particle-mesh Ewald (PME) or a reaction field (RF) were compared. Structural properties such as the area per lipid, carbon-deuterium order parameters, electron d. profile and bilayer thicknesses, are reproduced by all the parameter sets within the uncertainty of the available exptl. data. However, there are clear differences in the ordering of the glycerol backbone and choline headgroup, and the orientation of the headgroup dipole. In some cases, the degree of ordering was reminiscent of a liq.-ordered phase. It is also shown that, although the lateral diffusion of the lipids in the plane of the bilayer is often used to validate lipid force fields, because of the uncertainty in the exptl. measurements and the fact that the lateral diffusion is dependent on the choice of the simulation conditions, it should not be employed as a measure of quality. Finally, the simulations show that the effect of small changes in force-field parameters on the structure and dynamics of a bilayer is more significant than the treatment of the long-range electrostatic interactions using RF or PME. Overall, the GROMOS 54A7 best reproduced the range of exptl. data examd. - 106Wohlert, J.; Edholm, O. Dynamics in atomistic simulations of phospholipid membranes: Nuclear magnetic resonance relaxation rates and lateral diffusion J. Chem. Phys. 2006, 125 (20) 204703
- 107Wang, Y.; Markwick, P. R. L.; de Oliveira, C. A. F.; McCammon, J. A. Enhanced lipid diffusion and mixing in accelerated molecular dynamics J. Chem. Theory Comput. 2011, 7 (10) 3199– 3207[ACS Full Text
], [CAS], Google Scholar
107https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFCmtL7K&md5=e170125dc7540beccc910d1d542a4247Enhanced Lipid Diffusion and Mixing in Accelerated Molecular DynamicsWang, Yi; Markwick, Phineus R. L.; de Oliveira, Cesar Augusto F.; McCammon, J. AndrewJournal of Chemical Theory and Computation (2011), 7 (10), 3199-3207CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Accelerated mol. dynamics (aMD) is an enhanced sampling technique that expedites conformational space sampling by reducing the barriers sepg. various low-energy states of a system. Here, the authors present the first application of the aMD method on lipid membranes. Altogether, ∼1.5 μs simulations were performed on three systems: a pure POPC bilayer, a pure DMPC bilayer, and a mixed POPC:DMPC bilayer. Overall, the aMD simulations are found to produce significant speedup in trans-gauche isomerization and lipid lateral diffusion vs. those in conventional MD (cMD) simulations. Further comparison of a 70-ns aMD run and a 300-ns cMD run of the mixed POPC:DMPC bilayer shows that the two simulations yield similar lipid mixing behaviors, with aMD generating a 2-3-fold speedup compared to cMD. The authors' results demonstrate that the aMD method is an efficient approach for the study of bilayer structural and dynamic properties. On the basis of simulations of the three bilayer systems, the authors also discuss the impact of aMD parameters on various lipid properties, which can be used as a guideline for future aMD simulations of membrane systems. - 108Basconi, J. E.; Shirts, M. R. Effects of temperature control algorithms on transport properties and kinetics in molecular dynamics simulations J. Chem. Theory Comput. 2013, 9 (7) 2887– 2899[ACS Full Text
], [CAS], Google Scholar
108https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXnvVKhsbg%253D&md5=d3cd49b892dc363f4177d8eca8aaad34Effects of Temperature Control Algorithms on Transport Properties and Kinetics in Molecular Dynamics SimulationsBasconi, Joseph E.; Shirts, Michael R.Journal of Chemical Theory and Computation (2013), 9 (7), 2887-2899CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Temp. control algorithms in mol. dynamics (MD) simulations are necessary to study isothermal systems. However, these thermostatting algorithms alter the velocities of the particles and thus modify the dynamics of the system with respect to the microcanonical ensemble, which could potentially lead to thermostat-dependent dynamical artifacts. In this study, we investigate how six well-established thermostat algorithms applied with different coupling strengths and to different degrees of freedom affect the dynamics of various mol. systems. We consider dynamic processes occurring on different times scales by measuring translational and rotational self-diffusion as well as the shear viscosity of water, diffusion of a small mol. solvated in water, and diffusion and the dynamic structure factor of a polymer chain in water. All of these properties are significantly dampened by thermostat algorithms which randomize particle velocities, such as the Andersen thermostat and Langevin dynamics, when strong coupling is used. For the solvated small mol. and polymer, these dampening effects are reduced somewhat if the thermostats are applied to the solvent alone, such that the solute's temp. is maintained only through thermal contact with solvent particles. Algorithms which operate by scaling the velocities, such as the Berendsen thermostat, the stochastic velocity rescaling approach of Bussi and co-workers, and the Nose-Hoover thermostat, yield transport properties that are statistically indistinguishable from those of the microcanonical ensemble, provided they are applied globally, i.e. coupled to the system's kinetic energy. When coupled to local kinetic energies, a velocity scaling thermostat can have dampening effects comparable to a velocity randomizing method, as we observe when a massive Nose-Hoover coupling scheme is used to simulate water. Correct dynamical properties, at least those studied in this paper, are obtained with the Berendsen thermostat applied globally, despite the fact that it yields the wrong kinetic energy distribution. - 109König, S.; Bayerl, T. M.; Coddens, G.; Richter, D.; Sackmann, E. Hydration dependence of chain dynamics and local diffusion in L-alpha-dipalmitoylphosphtidylcholine multilayers studied by incoherent quasi-elastic neutron scattering Biophys. J. 1995, 68 (5) 1871– 1880
Cited By
This article is cited by 804 publications.
- Dulce C. Guzmán-Ocampo, Rodrigo Aguayo-Ortiz, José-Luis Velasco-Bolom, Pancham Lal Gupta, Adrian E. Roitberg, Laura Dominguez. Elucidating the Protonation State of the γ-Secretase Catalytic Dyad. ACS Chemical Neuroscience 2023, 14 (2) , 261-269. https://doi.org/10.1021/acschemneuro.2c00563
- Anastasiia Delova, Raúl Losantos, Jérémy Pecourneau, Yann Bernhard, Maxime Mourer, Andreea Pasc, Antonio Monari. Perturbation of Lipid Bilayers by Biomimetic Photoswitches Based on Cyclocurcumin. Journal of Chemical Information and Modeling 2023, 63 (1) , 299-307. https://doi.org/10.1021/acs.jcim.2c01152
- Ryuhei Harada, Rikuri Morita, Yasuteru Shigeta. Free-Energy Profiles for Membrane Permeation of Compounds Calculated Using Rare-Event Sampling Methods. Journal of Chemical Information and Modeling 2023, 63 (1) , 259-269. https://doi.org/10.1021/acs.jcim.2c01097
- Milla Kurki, Antti Poso, Piia Bartos, Markus S. Miettinen. Structure of POPC Lipid Bilayers in OPLS3e Force Field. Journal of Chemical Information and Modeling 2022, 62 (24) , 6462-6474. https://doi.org/10.1021/acs.jcim.2c00395
- Jérémy Pecourneau, Raúl Losantos, Anastasiia Delova, Yann Bernhard, Stéphane Parant, Maxime Mourer, Antonio Monari, Andreea Pasc. Biomimetic Photo-Switches Softening Model Lipid Membranes. Langmuir 2022, 38 (50) , 15642-15655. https://doi.org/10.1021/acs.langmuir.2c02425
- An Xiaoli, Niu Yuzhen, Yang Qiong, Lei Yang, Xiaojun Yao, Zhitong Bing. Investigating the Dynamic Binding Behavior of PMX53 Cooperating with Allosteric Antagonist NDT9513727 to C5a Anaphylatoxin Chemotactic Receptor 1 through Gaussian Accelerated Molecular Dynamics and Free-Energy Perturbation Simulations. ACS Chemical Neuroscience 2022, 13 (23) , 3502-3511. https://doi.org/10.1021/acschemneuro.2c00556
- Nil Casajuana-Martin, Gemma Navarro, Angel Gonzalez, Claudia Llinas del Torrent, Marc Gómez-Autet, Aleix Quintana García, Rafael Franco, Leonardo Pardo. A Single Point Mutation Blocks the Entrance of Ligands to the Cannabinoid CB2 Receptor via the Lipid Bilayer. Journal of Chemical Information and Modeling 2022, 62 (22) , 5771-5779. https://doi.org/10.1021/acs.jcim.2c00865
- Tingting Fu, Hongxing Zhang, Qingchuan Zheng. Molecular Insights into the Heterotropic Allosteric Mechanism in Cytochrome P450 3A4-Mediated Midazolam Metabolism. Journal of Chemical Information and Modeling 2022, 62 (22) , 5762-5770. https://doi.org/10.1021/acs.jcim.2c01264
- Mengrong Li, Miaomiao Li, Jingjing Guo. Molecular Mechanism of Ca2+ in the Allosteric Regulation of Human Parathyroid Hormone Receptor-1. Journal of Chemical Information and Modeling 2022, 62 (21) , 5110-5119. https://doi.org/10.1021/acs.jcim.1c00471
- Dongxiao Hao, He Wang, Yongjian Zang, Lei Zhang, Zhiwei Yang, Shengli Zhang. Mechanism of Glycans Modulating Cholesteryl Ester Transfer Protein: Unveiled by Molecular Dynamics Simulation. Journal of Chemical Information and Modeling 2022, 62 (21) , 5246-5257. https://doi.org/10.1021/acs.jcim.1c00233
- Fuhui Zhang, Xin Chen, Jianfang Chen, Yanjiani Xu, Shiqi Li, Yanzhi Guo, Xuemei Pu. Probing Allosteric Regulation Mechanism of W7.35 on Agonist-Induced Activity for μOR by Mutation Simulation. Journal of Chemical Information and Modeling 2022, 62 (21) , 5120-5135. https://doi.org/10.1021/acs.jcim.1c00650
- Humanath Poudel, David M. Leitner. Energy Transport in Class B GPCRs: Role of Protein–Water Dynamics and Activation. The Journal of Physical Chemistry B 2022, 126 (42) , 8362-8373. https://doi.org/10.1021/acs.jpcb.2c03960
- Lukács J. Németh, Tamás A. Martinek, Balázs Jójárt. Tilted State Population of Antimicrobial Peptide PGLa Is Coupled to the Transmembrane Potential. Journal of Chemical Information and Modeling 2022, 62 (20) , 4963-4969. https://doi.org/10.1021/acs.jcim.2c00667
- Adam Stasiulewicz, Anna Lesniak, Piotr Setny, Magdalena Bujalska-Zadrożny, Joanna I. Sulkowska. Identification of CB1 Ligands among Drugs, Phytochemicals and Natural-Like Compounds: Virtual Screening and In Vitro Verification. ACS Chemical Neuroscience 2022, 13 (20) , 2991-3007. https://doi.org/10.1021/acschemneuro.2c00502
- Tianyi Ding, Dmitry S. Karlov, Almudena Pino-Angeles, Irina G. Tikhonova. Intermolecular Interactions in G Protein-Coupled Receptor Allosteric Sites at the Membrane Interface from Molecular Dynamics Simulations and Quantum Chemical Calculations. Journal of Chemical Information and Modeling 2022, 62 (19) , 4736-4747. https://doi.org/10.1021/acs.jcim.2c00788
- Anastasia Croitoru, Alexey Aleksandrov. Parametrization of Force Field Bonded Terms under Structural Inconsistency. Journal of Chemical Information and Modeling 2022, 62 (19) , 4771-4782. https://doi.org/10.1021/acs.jcim.2c00950
- Javier García-Cárceles, Henar Vázquez-Villa, José Brea, David Ladron de Guevara-Miranda, Giovanni Cincilla, Melchor Sánchez-Martínez, Anabel Sánchez-Merino, Sergio Algar, María Teresa de los Frailes, Richard S. Roberts, Juan A. Ballesteros, Fernando Rodríguez de Fonseca, Bellinda Benhamú, María I. Loza, María L. López-Rodríguez. 2-(Fluoromethoxy)-4′-(S-methanesulfonimidoyl)-1,1′-biphenyl (UCM-1306), an Orally Bioavailable Positive Allosteric Modulator of the Human Dopamine D1 Receptor for Parkinson’s Disease. Journal of Medicinal Chemistry 2022, 65 (18) , 12256-12272. https://doi.org/10.1021/acs.jmedchem.2c00949
- Arslan R. Shaimardanov, Dmitry A. Shulga, Vladimir A. Palyulin. Is an Inductive Effect Explicit Account Required for Atomic Charges Aimed at Use within the Force Fields?. The Journal of Physical Chemistry A 2022, 126 (36) , 6278-6294. https://doi.org/10.1021/acs.jpca.2c02722
- Lucile Moynié, Françoise Hoegy, Stefan Milenkovic, Mathilde Munier, Aurélie Paulen, Véronique Gasser, Aline L. Faucon, Nicolas Zill, James H. Naismith, Matteo Ceccarelli, Isabelle J. Schalk, Gaëtan L.A. Mislin. Hijacking of the Enterobactin Pathway by a Synthetic Catechol Vector Designed for Oxazolidinone Antibiotic Delivery in Pseudomonas aeruginosa. ACS Infectious Diseases 2022, 8 (9) , 1894-1904. https://doi.org/10.1021/acsinfecdis.2c00202
- Anjana P. Menon, Wanqian Dong, Tzong-Hsien Lee, Marie-Isabel Aguilar, Mojie Duan, Shobhna Kapoor. Mutually Exclusive Interactions of Rifabutin with Spatially Distinct Mycobacterial Cell Envelope Membrane Layers Offer Insights into Membrane-Centric Therapy of Infectious Diseases. ACS Bio & Med Chem Au 2022, 2 (4) , 395-408. https://doi.org/10.1021/acsbiomedchemau.2c00010
- Shifan Ma, Zhuyezi Sun, Yankang Jing, Mark McGann, Sandor Vajda, Istvan J. Enyedy. Use of Solvent Mapping for Characterizing the Binding Site and for Predicting the Inhibition of the Human Ether-á-Go-Go-Related K+ Channel. Chemical Research in Toxicology 2022, 35 (8) , 1359-1369. https://doi.org/10.1021/acs.chemrestox.2c00036
- Laust Moesgaard, Peter Reinholdt, Carsten Uhd Nielsen, Jacob Kongsted. Mechanism behind Polysorbates’ Inhibitory Effect on P-Glycoprotein. Molecular Pharmaceutics 2022, 19 (7) , 2248-2253. https://doi.org/10.1021/acs.molpharmaceut.2c00074
- Yuna Lee, Akihiro Nakano, Yuki Nagasato, Takashi Ichinose, Toshiro Matsui. In Vitro and in Silico Analyses of the Adiponectin Receptor Agonistic Action of Soybean Tripeptides. Journal of Agricultural and Food Chemistry 2022, 70 (25) , 7695-7703. https://doi.org/10.1021/acs.jafc.2c02115
- Yoshiki Ishii, Nobuyuki Matubayasi, Hitoshi Washizu. Nonpolarizable Force Fields through the Self-Consistent Modeling Scheme with MD and DFT Methods: From Ionic Liquids to Self-Assembled Ionic Liquid Crystals. The Journal of Physical Chemistry B 2022, 126 (24) , 4611-4622. https://doi.org/10.1021/acs.jpcb.2c02782
- Mengrong Li, Yiqiong Bao, Ran Xu, Honggui La, Jingjing Guo. Critical Extracellular Ca2+ Dependence of the Binding between PTH1R and a G-Protein Peptide Revealed by MD Simulations. ACS Chemical Neuroscience 2022, 13 (11) , 1666-1674. https://doi.org/10.1021/acschemneuro.2c00176
- She Zhang, Jeff P. Thompson, Junchao Xia, Anthony T. Bogetti, Forrest York, A. Geoffrey Skillman, Lillian T. Chong, David N. LeBard. Mechanistic Insights into Passive Membrane Permeability of Drug-like Molecules from a Weighted Ensemble of Trajectories. Journal of Chemical Information and Modeling 2022, 62 (8) , 1891-1904. https://doi.org/10.1021/acs.jcim.1c01540
- Gaurav Sharma, Kenneth M. Merz. Mechanism of Zinc Transport through the Zinc Transporter YiiP. Journal of Chemical Theory and Computation 2022, 18 (4) , 2556-2568. https://doi.org/10.1021/acs.jctc.1c00927
- Chris J. Malajczuk, Sławomir S. Stachura, James O. Hendry, Ricardo L. Mancera. Redefining the Molecular Interplay between Dimethyl Sulfoxide, Lipid Bilayers, and Dehydration. The Journal of Physical Chemistry B 2022, 126 (13) , 2513-2529. https://doi.org/10.1021/acs.jpcb.2c00353
- Ruibin Liang, Amirhossein Bakhtiiari. Effects of Enzyme–Ligand Interactions on the Photoisomerization of a Light-Regulated Chemotherapeutic Drug. The Journal of Physical Chemistry B 2022, 126 (12) , 2382-2393. https://doi.org/10.1021/acs.jpcb.1c10819
- Callum J. Dickson, Ross C. Walker, Ian R. Gould. Lipid21: Complex Lipid Membrane Simulations with AMBER. Journal of Chemical Theory and Computation 2022, 18 (3) , 1726-1736. https://doi.org/10.1021/acs.jctc.1c01217
- James Andrews, Estela Blaisten-Barojas. Distinctive Formation of PEG-Lipid Nanopatches onto Solid Polymer Surfaces Interfacing Solvents from Atomistic Simulation. The Journal of Physical Chemistry B 2022, 126 (7) , 1598-1608. https://doi.org/10.1021/acs.jpcb.1c07490
- Weiwei Xue, Tingting Fu, Shengzhe Deng, Fengyuan Yang, Jingyi Yang, Feng Zhu. Molecular Mechanism for the Allosteric Inhibition of the Human Serotonin Transporter by Antidepressant Escitalopram. ACS Chemical Neuroscience 2022, 13 (3) , 340-351. https://doi.org/10.1021/acschemneuro.1c00694
- Chris J. Malajczuk, Blake I. Armstrong, Sławomir S. Stachura, Ricardo L. Mancera. Mechanisms of Interaction of Small Hydroxylated Cryosolvents with Dehydrated Model Cell Membranes: Stabilization vs Destruction. The Journal of Physical Chemistry B 2022, 126 (1) , 197-216. https://doi.org/10.1021/acs.jpcb.1c07769
- Daniel Pulido, Verònica Casadó-Anguera, Marc Gómez-Autet, Natàlia Llopart, Estefanía Moreno, Nil Casajuana-Martin, Sergi Ferré, Leonardo Pardo, Vicent Casadó, Miriam Royo. Heterobivalent Ligand for the Adenosine A2A–Dopamine D2 Receptor Heteromer. Journal of Medicinal Chemistry 2022, 65 (1) , 616-632. https://doi.org/10.1021/acs.jmedchem.1c01763
- Min-Kang Hsieh, Yalun Yu, Jeffery B. Klauda. All-Atom Modeling of Complex Cellular Membranes. Langmuir 2022, 38 (1) , 3-17. https://doi.org/10.1021/acs.langmuir.1c02084
- Callum J Dickson, Viktor Hornak, Jose S Duca. Relative Binding Free-Energy Calculations at Lipid-Exposed Sites: Deciphering Hot Spots. Journal of Chemical Information and Modeling 2021, 61 (12) , 5923-5930. https://doi.org/10.1021/acs.jcim.1c01147
- Satoshi Ono, Matthew R. Naylor, Chad E. Townsend, Chieko Okumura, Okimasa Okada, Hsiau-Wei Lee, R. Scott Lokey. Cyclosporin A: Conformational Complexity and Chameleonicity. Journal of Chemical Information and Modeling 2021, 61 (11) , 5601-5613. https://doi.org/10.1021/acs.jcim.1c00771
- Sabahuddin Ahmad, Christoph Heinrich Strunk, Stephan N. Schott-Verdugo, Karl-Erich Jaeger, Filip Kovacic, Holger Gohlke. Substrate Access Mechanism in a Novel Membrane-Bound Phospholipase A of Pseudomonas aeruginosa Concordant with Specificity and Regioselectivity. Journal of Chemical Information and Modeling 2021, 61 (11) , 5626-5643. https://doi.org/10.1021/acs.jcim.1c00973
- Gert-Jan Bekker, Mitsugu Araki, Kanji Oshima, Yasushi Okuno, Narutoshi Kamiya. Accurate Binding Configuration Prediction of a G-Protein-Coupled Receptor to Its Antagonist Using Multicanonical Molecular Dynamics-Based Dynamic Docking. Journal of Chemical Information and Modeling 2021, 61 (10) , 5161-5171. https://doi.org/10.1021/acs.jcim.1c00712
- Daria B. Kokh, Rebecca C. Wade. G Protein-Coupled Receptor–Ligand Dissociation Rates and Mechanisms from τRAMD Simulations. Journal of Chemical Theory and Computation 2021, 17 (10) , 6610-6623. https://doi.org/10.1021/acs.jctc.1c00641
- Laura Jenkins, Sara Marsango, Sarah Mancini, Zobaer Al Mahmud, Angus Morrison, Stuart P. McElroy, Kirstie A. Bennett, Matt Barnes, Andrew B. Tobin, Irina G. Tikhonova, Graeme Milligan. Discovery and Characterization of Novel Antagonists of the Proinflammatory Orphan Receptor GPR84. ACS Pharmacology & Translational Science 2021, 4 (5) , 1598-1613. https://doi.org/10.1021/acsptsci.1c00151
- Hiromi Nakai, Toshiaki Takemura, Junichi Ono, Yoshifumi Nishimura. Quantum-Mechanical Molecular Dynamics Simulations on Secondary Proton Transfer in Bacteriorhodopsin Using Realistic Models. The Journal of Physical Chemistry B 2021, 125 (39) , 10947-10963. https://doi.org/10.1021/acs.jpcb.1c06231
- Kazuhiro J. Fujimoto, Takumi Minoda, Takeshi Yanai. Spectral Tuning Mechanism of Photosynthetic Light-Harvesting Complex II Revealed by Ab Initio Dimer Exciton Model. The Journal of Physical Chemistry B 2021, 125 (37) , 10459-10470. https://doi.org/10.1021/acs.jpcb.1c04457
- Gaurav Sharma, Kenneth M. Merz. Formation of the Metal-Binding Core of the ZRT/IRT-like Protein (ZIP) Family Zinc Transporter. Biochemistry 2021, 60 (36) , 2727-2738. https://doi.org/10.1021/acs.biochem.1c00415
- Tetsuo Komori, Heri Satria, Kyohei Miyamura, Ai Ito, Magoto Kamiya, Ayumi Sumino, Takakazu Onishi, Kazuaki Ninomiya, Kenji Takahashi, Jared L. Anderson, Takuya Uto, Kosuke Kuroda. Essential Requirements of Biocompatible Cellulose Solvents. ACS Sustainable Chemistry & Engineering 2021, 9 (35) , 11825-11836. https://doi.org/10.1021/acssuschemeng.1c03438
- Daniel Becker, Prasad V. Bharatam, Holger Gohlke. F/G Region Rigidity is Inversely Correlated to Substrate Promiscuity of Human CYP Isoforms Involved in Metabolism. Journal of Chemical Information and Modeling 2021, 61 (8) , 4023-4030. https://doi.org/10.1021/acs.jcim.1c00558
- Christopher Faulkner, Nora H. de Leeuw. Predicting the Membrane Permeability of Fentanyl and Its Analogues by Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2021, 125 (30) , 8443-8449. https://doi.org/10.1021/acs.jpcb.1c05438
- Pavel Janoš, Alessandra Magistrato. All-Atom Simulations Uncover the Molecular Terms of the NKCC1 Transport Mechanism. Journal of Chemical Information and Modeling 2021, 61 (7) , 3649-3658. https://doi.org/10.1021/acs.jcim.1c00551
- Olaia Martí-Marí, Belén Martínez-Gualda, Sofía de la Puente-Secades, Alberto Mills, Ernesto Quesada, Rana Abdelnabi, Liang Sun, Arnaud Boonen, Sam Noppen, Johan Neyts, Dominique Schols, María-José Camarasa, Federico Gago, Ana San-Félix. Double Arylation of the Indole Side Chain of Tri- and Tetrapodal Tryptophan Derivatives Renders Highly Potent HIV-1 and EV-A71 Entry Inhibitors. Journal of Medicinal Chemistry 2021, 64 (14) , 10027-10046. https://doi.org/10.1021/acs.jmedchem.1c00315
- Thomas Durek, Quentin Kaas, Andrew M. White, Joachim Weidmann, Abdullah Ahmad Fuaad, Olivier Cheneval, Christina I. Schroeder, Simon J. de Veer, Anita Dellsén, Torben Österlund, Niklas Larsson, Laurent Knerr, Udo Bauer, Alleyn T. Plowright, David J. Craik. Melanocortin 1 Receptor Agonists Based on a Bivalent, Bicyclic Peptide Framework. Journal of Medicinal Chemistry 2021, 64 (14) , 9906-9915. https://doi.org/10.1021/acs.jmedchem.1c00095
- Humanath Poudel, David M. Leitner. Activation-Induced Reorganization of Energy Transport Networks in the β2 Adrenergic Receptor. The Journal of Physical Chemistry B 2021, 125 (24) , 6522-6531. https://doi.org/10.1021/acs.jpcb.1c03412
- Dehui Zhang, David A. Perrey, Ann M. Decker, Tiffany L. Langston, Vijayakumar Mavanji, Danni L. Harris, Catherine M. Kotz, Yanan Zhang. Discovery of Arylsulfonamides as Dual Orexin Receptor Agonists. Journal of Medicinal Chemistry 2021, 64 (12) , 8806-8825. https://doi.org/10.1021/acs.jmedchem.1c00841
- Jeffery B. Klauda. Considerations of Recent All-Atom Lipid Force Field Development. The Journal of Physical Chemistry B 2021, 125 (22) , 5676-5682. https://doi.org/10.1021/acs.jpcb.1c02417
- Goli Yamini, Subbarao Kanchi, Nnanya Kalu, Sanaz Momben Abolfath, Stephen H. Leppla, K. Ganapathy Ayappa, Prabal K. Maiti, Ekaterina M. Nestorovich. Hydrophobic Gating and 1/f Noise of the Anthrax Toxin Channel. The Journal of Physical Chemistry B 2021, 125 (21) , 5466-5478. https://doi.org/10.1021/acs.jpcb.0c10490
- Gao Tu, Tingting Fu, Fengyuan Yang, Jingyi Yang, Zhao Zhang, Xiaojun Yao, Weiwei Xue, Feng Zhu. Understanding the Polypharmacological Profiles of Triple Reuptake Inhibitors by Molecular Simulation. ACS Chemical Neuroscience 2021, 12 (11) , 2013-2026. https://doi.org/10.1021/acschemneuro.1c00127
- Junhao Li, Yue Chen, Yun Tang, Weihua Li, Yaoquan Tu. Homotropic Cooperativity of Midazolam Metabolism by Cytochrome P450 3A4: Insight from Computational Studies. Journal of Chemical Information and Modeling 2021, 61 (5) , 2418-2426. https://doi.org/10.1021/acs.jcim.1c00266
- Shun Yokoi, Ayori Mitsutake. Molecular Dynamics Simulations for the Determination of the Characteristic Structural Differences between Inactive and Active States of Wild Type and Mutants of the Orexin2 Receptor. The Journal of Physical Chemistry B 2021, 125 (17) , 4286-4298. https://doi.org/10.1021/acs.jpcb.0c10985
- Jin Cheng, Maozi Chen, Siyi Wang, Tianjian Liang, Hui Chen, Chih-Jung Chen, Zhiwei Feng, Xiang-Qun Xie. Binding Characterization of Agonists and Antagonists by MCCS: A Case Study from Adenosine A2A Receptor. ACS Chemical Neuroscience 2021, 12 (9) , 1606-1620. https://doi.org/10.1021/acschemneuro.1c00082
- Xinghang Yuan, Di Zhang, Shengjun Mao, Qiantao Wang. Filling the Gap in Understanding the Mechanism of GABAAR and Propofol Using Computational Approaches. Journal of Chemical Information and Modeling 2021, 61 (4) , 1889-1901. https://doi.org/10.1021/acs.jcim.0c01290
- Ruibin Liang, Jimmy K. Yu, Jan Meisner, Fang Liu, Todd J. Martinez. Electrostatic Control of Photoisomerization in Channelrhodopsin 2. Journal of the American Chemical Society 2021, 143 (14) , 5425-5437. https://doi.org/10.1021/jacs.1c00058
- Rafael Santana Nunes, Diogo Vila-Viçosa, Paulo J. Costa. Halogen Bonding: An Underestimated Player in Membrane–Ligand Interactions. Journal of the American Chemical Society 2021, 143 (11) , 4253-4267. https://doi.org/10.1021/jacs.0c12470
- Yalun Yu, Andreas Krämer, Richard M. Venable, Andrew C. Simmonett, Alexander D. MacKerell, Jr., Jeffery B. Klauda, Richard W. Pastor, Bernard R. Brooks. Semi-automated Optimization of the CHARMM36 Lipid Force Field to Include Explicit Treatment of Long-Range Dispersion. Journal of Chemical Theory and Computation 2021, 17 (3) , 1562-1580. https://doi.org/10.1021/acs.jctc.0c01326
- Abhishek Sirohiwal, Frank Neese, Dimitrios A. Pantazis. How Can We Predict Accurate Electrochromic Shifts for Biochromophores? A Case Study on the Photosynthetic Reaction Center. Journal of Chemical Theory and Computation 2021, 17 (3) , 1858-1873. https://doi.org/10.1021/acs.jctc.0c01152
- Brian H. Morrow, Judith A. Harrison. Evaluating the Ability of Selected Force Fields to Simulate Hydrocarbons as a Function of Temperature and Pressure Using Molecular Dynamics. Energy & Fuels 2021, 35 (5) , 3742-3752. https://doi.org/10.1021/acs.energyfuels.0c03363
- Jovan Damjanovic, Jiayuan Miao, He Huang, Yu-Shan Lin. Elucidating Solution Structures of Cyclic Peptides Using Molecular Dynamics Simulations. Chemical Reviews 2021, 121 (4) , 2292-2324. https://doi.org/10.1021/acs.chemrev.0c01087
- Hanne S. Antila, Tiago M. Ferreira, O. H. Samuli Ollila, Markus S. Miettinen. Using Open Data to Rapidly Benchmark Biomolecular Simulations: Phospholipid Conformational Dynamics. Journal of Chemical Information and Modeling 2021, 61 (2) , 938-949. https://doi.org/10.1021/acs.jcim.0c01299
- Shiji Zhao, Andrew J. Schaub, Shiou-Chuan Tsai, Ray Luo. Development of a Pantetheine Force Field Library for Molecular Modeling. Journal of Chemical Information and Modeling 2021, 61 (2) , 856-868. https://doi.org/10.1021/acs.jcim.0c01384
- Deborah Thomas, Vicente Rubio, Vijaya Iragavarapu, Esther Guzman, Oliver B. Pelletier, Shahriar Alamgir, Qi Zhang, Maciej J. Stawikowski. Solvatochromic and pH-Sensitive Fluorescent Membrane Probes for Imaging of Live Cells. ACS Chemical Neuroscience 2021, 12 (4) , 719-734. https://doi.org/10.1021/acschemneuro.0c00732
- Zack Jarin, James Newhouse, Gregory A. Voth. Coarse-Grained Force Fields from the Perspective of Statistical Mechanics: Better Understanding of the Origins of a MARTINI Hangover. Journal of Chemical Theory and Computation 2021, 17 (2) , 1170-1180. https://doi.org/10.1021/acs.jctc.0c00638
- Loknath Dhar, Saddam Hossain, Md Sajjadur Rahman, Shamshad B. Quraishi, Koushik Saha, Farzana Rahman, Mir Tamzid Rahman. Adsorption Mechanism of Methylene Blue by Graphene Oxide-Shielded Mg–Al-Layered Double Hydroxide From Synthetic Wastewater. The Journal of Physical Chemistry A 2021, 125 (4) , 954-965. https://doi.org/10.1021/acs.jpca.0c09124
- Siddhartha Banerjee, Mohtadin Hashemi, Karen Zagorski, Yuri L. Lyubchenko. Cholesterol in Membranes Facilitates Aggregation of Amyloid β Protein at Physiologically Relevant Concentrations. ACS Chemical Neuroscience 2021, 12 (3) , 506-516. https://doi.org/10.1021/acschemneuro.0c00688
- Ipsita Basu, Prabal K. Maiti. Insight into the Mechanism of Carrier-Mediated Delivery of siRNA in the Cell Membrane Using MD Simulation. Langmuir 2021, 37 (1) , 266-277. https://doi.org/10.1021/acs.langmuir.0c02871
- Matti Javanainen, Wei Hua, Ondrej Tichacek, Pauline Delcroix, Lukasz Cwiklik, Heather C. Allen. Structural Effects of Cation Binding to DPPC Monolayers. Langmuir 2020, 36 (50) , 15258-15269. https://doi.org/10.1021/acs.langmuir.0c02555
- Matheus Henrique Reis, Deborah Antunes, Lucianna H. S. Santos, Ana Carolina Ramos Guimarães, Ernesto Raul Caffarena. Shared Binding Mode of Perrottetinene and Tetrahydrocannabinol Diastereomers inside the CB1 Receptor May Incentivize Novel Medicinal Drug Design: Findings from an in Silico Assay. ACS Chemical Neuroscience 2020, 11 (24) , 4289-4300. https://doi.org/10.1021/acschemneuro.0c00547
- Mara Silber, Manuel Hitzenberger, Martin Zacharias, Claudia Muhle-Goll. Altered Hinge Conformations in APP Transmembrane Helix Mutants May Affect Enzyme–Substrate Interactions of γ-Secretase. ACS Chemical Neuroscience 2020, 11 (24) , 4426-4433. https://doi.org/10.1021/acschemneuro.0c00640
- Christine Robinson, Valeria Gradinati, Fatima Hamid, Carly Baehr, Bethany Crouse, Saadyah Averick, Marina Kovaliov, Danni Harris, Scott Runyon, Federico Baruffaldi, Mark LeSage, Sandra Comer, Marco Pravetoni. Therapeutic and Prophylactic Vaccines to Counteract Fentanyl Use Disorders and Toxicity. Journal of Medicinal Chemistry 2020, 63 (23) , 14647-14667. https://doi.org/10.1021/acs.jmedchem.0c01042
- Cecilie Søderlund Kofod, Salvatore Prioli, Mick Hornum, Jacob Kongsted, Peter Reinholdt. Computational Characterization of Novel Malononitrile Variants of Laurdan with Improved Photophysical Properties for Sensing in Membranes. The Journal of Physical Chemistry B 2020, 124 (43) , 9526-9534. https://doi.org/10.1021/acs.jpcb.0c06011
- Michael A. Johnston, Andrew Ian Duff, Richard L. Anderson, William C. Swope. Model for the Simulation of the CnEm Nonionic Surfactant Family Derived from Recent Experimental Results. The Journal of Physical Chemistry B 2020, 124 (43) , 9701-9721. https://doi.org/10.1021/acs.jpcb.0c06132
- Carmen Esposito, Shuzhe Wang, Udo E. W. Lange, Frank Oellien, Sereina Riniker. Combining Machine Learning and Molecular Dynamics to Predict P-Glycoprotein Substrates. Journal of Chemical Information and Modeling 2020, 60 (10) , 4730-4749. https://doi.org/10.1021/acs.jcim.0c00525
- Zhiwei Feng, Tianjian Liang, Siyi Wang, Maozi Chen, Tianling Hou, Jack Zhao, Hui Chen, Yuehan Zhou, Xiang-Qun Xie. Binding Characterization of GPCRs-Modulator by Molecular Complex Characterizing System (MCCS). ACS Chemical Neuroscience 2020, 11 (20) , 3333-3345. https://doi.org/10.1021/acschemneuro.0c00457
- Abhishek Sirohiwal, Frank Neese, Dimitrios A. Pantazis. Protein Matrix Control of Reaction Center Excitation in Photosystem II. Journal of the American Chemical Society 2020, 142 (42) , 18174-18190. https://doi.org/10.1021/jacs.0c08526
- Violetta Burns, Blake Mertz. Using Simulation to Understand the Role of Titration on the Stability of a Peptide–Lipid Bilayer Complex. Langmuir 2020, 36 (41) , 12272-12280. https://doi.org/10.1021/acs.langmuir.0c02038
- Didem Sanver, Amin Sadeghpour, Michael Rappolt, Florent Di Meo, Patrick Trouillas. Structure and Dynamics of Dioleoyl-Phosphatidylcholine Bilayers under the Influence of Quercetin and Rutin. Langmuir 2020, 36 (40) , 11776-11786. https://doi.org/10.1021/acs.langmuir.0c01484
- Alessio Prunotto, Guillermo Bahr, Lisandro J. González, Alejandro J. Vila, Matteo Dal Peraro. Molecular Bases of the Membrane Association Mechanism Potentiating Antibiotic Resistance by New Delhi Metallo-β-lactamase 1. ACS Infectious Diseases 2020, 6 (10) , 2719-2731. https://doi.org/10.1021/acsinfecdis.0c00341
- Hyeonuk Woo, Sang-Jun Park, Yeol Kyo Choi, Taeyong Park, Maham Tanveer, Yiwei Cao, Nathan R. Kern, Jumin Lee, Min Sun Yeom, Tristan I. Croll, Chaok Seok, Wonpil Im. Developing a Fully Glycosylated Full-Length SARS-CoV-2 Spike Protein Model in a Viral Membrane. The Journal of Physical Chemistry B 2020, 124 (33) , 7128-7137. https://doi.org/10.1021/acs.jpcb.0c04553
- Mounika Gosika, Vasumathi Velachi, M. Natália D. S. Cordeiro, Prabal K. Maiti. Covalent Functionalization of Graphene with PAMAM Dendrimer and Its Implications on Graphene’s Dispersion and Cytotoxicity. ACS Applied Polymer Materials 2020, 2 (8) , 3587-3600. https://doi.org/10.1021/acsapm.0c00596
- Conner M. Winkeljohn, Benjamin Himberg, Juan M. Vanegas. Balance of Solvent and Chain Interactions Determines the Local Stress State of Simulated Membranes. The Journal of Physical Chemistry B 2020, 124 (32) , 6963-6971. https://doi.org/10.1021/acs.jpcb.0c03937
- Yue Zhang, Hong-Xing Zhang, Qing-Chuan Zheng. In Silico Study of Membrane Lipid Composition Regulating Conformation and Hydration of Influenza Virus B M2 Channel. Journal of Chemical Information and Modeling 2020, 60 (7) , 3603-3615. https://doi.org/10.1021/acs.jcim.0c00329
- Jayesh Arun Bafna, Eulàlia Sans-Serramitjana, Silvia Acosta-Gutiérrez, Igor V. Bodrenko, Daniel Hörömpöli, Anne Berscheid, Heike Brötz-Oesterhelt, Mathias Winterhalter, Matteo Ceccarelli. Kanamycin Uptake into Escherichia coli Is Facilitated by OmpF and OmpC Porin Channels Located in the Outer Membrane. ACS Infectious Diseases 2020, 6 (7) , 1855-1865. https://doi.org/10.1021/acsinfecdis.0c00102
- Christopher Faulkner, David Santos-Carballal, David F. Plant, Nora H. de Leeuw. Atomistic Molecular Dynamics Simulations of Propofol and Fentanyl in Phosphatidylcholine Lipid Bilayers. ACS Omega 2020, 5 (24) , 14340-14353. https://doi.org/10.1021/acsomega.0c00813
- Pallavi Banerjee, Reinhard Lipowsky, Mark Santer. Coarse-Grained Molecular Model for the Glycosylphosphatidylinositol Anchor with and without Protein. Journal of Chemical Theory and Computation 2020, 16 (6) , 3889-3903. https://doi.org/10.1021/acs.jctc.0c00056
- Nozomu Kamiya, Megumi Kayanuma, Hideaki Fujitani, Keiko Shinoda. A New Lipid Force Field (FUJI). Journal of Chemical Theory and Computation 2020, 16 (6) , 3664-3676. https://doi.org/10.1021/acs.jctc.9b01195
- Tobias Klein, Frances D. Lenahan, Manuel Kerscher, Michael H. Rausch, Ioannis G. Economou, Thomas M. Koller, Andreas P. Fröba. Characterization of Long Linear and Branched Alkanes and Alcohols for Temperatures up to 573.15 K by Surface Light Scattering and Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2020, 124 (20) , 4146-4163. https://doi.org/10.1021/acs.jpcb.0c01740
- Christopher Lockhart, Amy K. Smith, Dmitri K. Klimov. Three Popular Force Fields Predict Consensus Mechanism of Amyloid β Peptide Binding to the Dimyristoylgylcerophosphocholine Bilayer. Journal of Chemical Information and Modeling 2020, 60 (4) , 2282-2293. https://doi.org/10.1021/acs.jcim.0c00096
- Dinh Quoc Huy Pham, Pawel Krupa, Hoang Linh Nguyen, Giovanni La Penna, Mai Suan Li. Computational Model to Unravel the Function of Amyloid-β Peptides in Contact with a Phospholipid Membrane. The Journal of Physical Chemistry B 2020, 124 (16) , 3300-3314. https://doi.org/10.1021/acs.jpcb.0c00771
- Beihong Ji, Shuhan Liu, Xibing He, Viet Hoang Man, Xiang-Qun Xie, Junmei Wang. Prediction of the Binding Affinities and Selectivity for CB1 and CB2 Ligands Using Homology Modeling, Molecular Docking, Molecular Dynamics Simulations, and MM-PBSA Binding Free Energy Calculations. ACS Chemical Neuroscience 2020, 11 (8) , 1139-1158. https://doi.org/10.1021/acschemneuro.9b00696
- Rilei Yu, Jiayi Wang, Lok-Yan So, Peta J. Harvey, Juan Shi, Jiazhen Liang, Qin Dou, Xiao Li, Xiayi Yan, Yen-Hua Huang, Qingliang Xu, Quentin Kaas, Ho-Yin Chow, Kwok-Yin Wong, David J. Craik, Xiao-Hua Zhang, Tao Jiang, Yan Wang. Enhanced Activity against Multidrug-Resistant Bacteria through Coapplication of an Analogue of Tachyplesin I and an Inhibitor of the QseC/B Signaling Pathway. Journal of Medicinal Chemistry 2020, 63 (7) , 3475-3484. https://doi.org/10.1021/acs.jmedchem.9b01563
- Franccesca Fornasier, Lucas M. P. Souza, Felipe R. Souza, Franceline Reynaud, Andre S. Pimentel. Lipophilicity of Coarse-Grained Cholesterol Models. Journal of Chemical Information and Modeling 2020, 60 (2) , 569-577. https://doi.org/10.1021/acs.jcim.9b00830
- Vanesa Racigh, Agustín Ormazábal, Juliana Palma, Gustavo Pierdominici-Sottile. Positively Charged Residues in the Head Domain of P2X4 Receptors Assist the Binding of ATP. Journal of Chemical Information and Modeling 2020, 60 (2) , 923-932. https://doi.org/10.1021/acs.jcim.9b00856
- Anita Wnętrzak, Anna Chachaj-Brekiesz, Jan Kobierski, Katarzyna Karwowska, Aneta D. Petelska, Patrycja Dynarowicz-Latka. Unusual Behavior of the Bipolar Molecule 25-Hydroxycholesterol at the Air/Water Interface—Langmuir Monolayer Approach Complemented with Theoretical Calculations. The Journal of Physical Chemistry B 2020, 124 (6) , 1104-1114. https://doi.org/10.1021/acs.jpcb.9b10938
Abstract
Figure 1
Figure 1. A box of 144 pentadecane molecules simulated in the NPT ensemble at 298.15 K using the General Amber Force Field (16) to model the carbon chains.
Figure 2
Figure 2. The energy profile for rotating about selected torsions of a cis-5-decene molecule. Energy evaluated using QM and the HM-IE method (filled triangle ▲), AMBER with standard GAFF parameters (dotted line), and AMBER with Lipid14 parameters (black line). Torsion fits from the top are as follows: CH2–CH–CH–CH2, CH–CH–CH2–CH2, and CH–CH2–CH2–CH2.
Figure 3
Figure 3. Structure and charges of Lipid11/Lipid14 headgroup and tail group caps. (22)
Figure 4
Figure 4. A capped lauroyl tail group residue was used to fit the oS-cC-cD-cD and oC-cC-cD-cD torsions.
Figure 5
Figure 5. The energy profiles for rotating about selected torsions of a capped lauroyl tail group residue. Energy evaluated using QM and the HM-IE method (filled triangle ▲), AMBER with standard GAFF/Lipid11 parameters (dotted line), and AMBER with Lipid14 parameters (black line). Torsion fits from the top are oC-cC-cD-cD and oS-cC-cD-cD.
Figure 6
Figure 6. Calculated 13C NMR T1 relaxation times for selected alkane chains and comparison to experiment. (47) Values at 312 K.
Figure 7
Figure 7. Simulation NMR order parameters for the six lipid systems and comparison to experiment. (77, 78, 80-84)
Figure 8
Figure 8. The total and decomposed electron density profiles for each of the six lipid bilayer systems with contributions from water, choline (CHOL), phosphate (PO4), glycerol (GLY), carbonyl (COO), methylene (CH2), unsaturated CH═CH and terminal methyls (CH3).
Figure 9
Figure 9. Simulation X-ray scattering form factors for the six lipid systems (black line) and comparison to experiment (54, 55, 62, 66, 68) (cyan circles). Inset: Simulation neutron scattering form factors at 100% D2O (black line), 70% D2O (red line), and 50% D2O (blue line) and comparison to experiment (55, 96) (black, red, and blue circles, respectively).
Figure 10
Figure 10. Plot of ΔDB-H versus area per lipid AL for the three all-atom lipid force fields CHARMM36 (squares), Slipids (diamonds), and AMBER Lipid14 (circles). Values shown for DLPC (green), DMPC (magenta), DPPC (blue), DOPC (red), and POPC (orange).
Figure 11
Figure 11. Time averaged mean square displacement of the center of mass of the lipid molecules versus NVE simulation time.
Figure 12
Figure 12. Lateral diffusion coefficients for the six lipid types calculated using different time ranges of the mean square displacement curve for the linear fit.
References
ARTICLE SECTIONSThis article references 109 other publications.
- 1van Meer, G.; Voelker, D. R.; Feigenson, G. W. Membrane lipids: where they are and how they behave Nat. Rev. Mol. Cell Biol. 2008, 9 (2) 112– 124[Crossref], [PubMed], [CAS], Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXovFOntw%253D%253D&md5=13d4eb41956211e22d3a9ee5c6b73854Membrane lipids: where they are and how they behavevan Meer, Gerrit; Voelker, Dennis R.; Feigenson, Gerald W.Nature Reviews Molecular Cell Biology (2008), 9 (2), 112-124CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. Throughout the biol. world, a 30 Å hydrophobic film typically delimits the environments that serve as the margin between life and death for individual cells. Biochem. and biophys. findings have provided a detailed model of the compn. and structure of membranes, which includes levels of dynamic organization both across the lipid bilayer (lipid asymmetry) and in the lateral dimension (lipid domains) of membranes. How do cells apply anabolic and catabolic enzymes, translocases, and transporters, plus the intrinsic phys. phase behavior of lipids and their interactions with membrane proteins, to create the unique compns. and multiple functionalities of their individual membranes.
- 2Lodish, H.; Berk, A.; Kaiser, C. A.; Scott, M. P.; Bretscher, A.; Ploegh, H.; Matsudaira, P. Molecular Cell Biology,6th ed.; W. H. Freeman: New York, 2007.Google ScholarThere is no corresponding record for this reference.
- 3Phillips, R.; Ursell, T.; Wiggins, P.; Sens, P. Emerging roles for lipids in shaping membrane-protein function Nature 2009, 459 (7245) 379– 385[Crossref], [PubMed], [CAS], Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmtFKmtrc%253D&md5=802acc0d4cf9ce32bd821d46f736e1ccEmerging roles for lipids in shaping membrane-protein functionPhillips, Rob; Ursell, Tristan; Wiggins, Paul; Sens, PierreNature (London, United Kingdom) (2009), 459 (7245), 379-385CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. Studies of membrane proteins have revealed a direct link between the lipid environment and the structure and function of some of these proteins. Although some of these effects involve specific chem. interactions between lipids and protein residues, many can be understood in terms of protein-induced perturbations to the membrane shape. The free-energy cost of such perturbations can be estd. quant., and measurements of channel gating in model systems of membrane proteins with their lipid partners are now confirming predictions of simple models.
- 4Nagle, J. F.; Tristram-Nagle, S. Structure of lipid bilayers Biochim. Biophys. Acta 2000, 1469 (3) 159– 195
- 5Katsaras, J.; Gutberlet, T. Lipid bilayers: Structure and interactions; Springer-Verlag: Berlin, 2001.
- 6Tieleman, D. P.; Marrink, S. J.; Berendsen, H. J. A computer perspective of membranes: molecular dynamics studies of lipid bilayer systems Biochim. Biophys. Acta 1997, 1331, 235– 270[Crossref], [PubMed], [CAS], Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXnvVSnsLw%253D&md5=58cecf9cb3f6c4d20d5a717eefbc4698A computer perspective of membranes: molecular dynamics studies of lipid bilayer systemsTieleman, D. P.; Marrink, S. J.; Berendsen, H. J. C.Biochimica et Biophysica Acta, Reviews on Biomembranes (1997), 1331 (3), 235-270CODEN: BRBMC5; ISSN:0304-4157. (Elsevier B.V.)A review with 125 refs. on mol. dynamic simulations and exptl. data which can be used to validate the simulations of lipid bilayers.
- 7Berger, O.; Edholm, O.; Jähnig, F. Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperature Biophys. J. 1997, 72, 2002– 2013[Crossref], [PubMed], [CAS], Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXivVOjsL4%253D&md5=984ea49ced9d0e2ce912bf99220fd228Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperatureBerger, Oliver; Edholm, Olle; Jahnig, FritzBiophysical Journal (1997), 72 (5), 2002-2013CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Mol. dynamics simulations of 500 ps were performed on a system consisting of a bilayer of 64 mols. of the lipid dipalmitoylphosphatidylcholine and 23 water mols. per lipid at an isotropic pressure of 1 atm and 50°. Special attention was devoted to reproduce the correct d. of the lipid, because this quantity is known exptl. with a precision better than 1%. For this purpose, the Lennard-Jones parameters of the hydrocarbon chains were adjusted by simulating a system consisting of 128 pentadecane mols. and varying the Lennard-Jones parameters until the exptl. d. and heat of vaporization were obtained. With these parameters the lipid d. resulted in perfect agreement with the exptl. d. The orientational order parameter of the hydrocarbon chains agreed perfectly well with the exptl. values, which, because of its correlation with the area per lipid, makes it possible to give a proper est. of the area per lipid of 0.61±0.1 nm2.
- 8Klauda, J. B.; Venable, R. M.; Freites, J. A.; O’Connor, J. W.; Tobias, D. J.; Mondragon-Ramirez, C.; Vorobyov, I.; MacKerell, A. D.; Pastor, R. W. Update of the CHARMM all-atom additive force field for lipids: Validation on Six lipid types J. Phys. Chem. B 2010, 114 (23) 7830– 7843[ACS Full Text
], [CAS], Google Scholar
8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXmsVGjtrk%253D&md5=04c4d3aa1a59740190994d7df80cbb71Update of the CHARMM All-Atom Additive Force Field for Lipids: Validation on Six Lipid TypesKlauda, Jeffery B.; Venable, Richard M.; Freites, J. Alfredo; O'Connor, Joseph W.; Tobias, Douglas J.; Mondragon-Ramirez, Carlos; Vorobyov, Igor; MacKerell, Alexander D., Jr.; Pastor, Richard W.Journal of Physical Chemistry B (2010), 114 (23), 7830-7843CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)A significant modification to the additive all-atom CHARMM lipid force field (FF) is developed and applied to phospholipid bilayers with both choline and ethanolamine contg. head groups and with both satd. and unsatd. aliph. chains. Motivated by the current CHARMM lipid FF (C27 and C27r) systematically yielding values of the surface area per lipid that are smaller than exptl. ests. and gel-like structures of bilayers well above the gel transition temp., selected torsional, Lennard-Jones and partial at. charge parameters were modified by targeting both quantum mech. (QM) and exptl. data. QM calcns. ranging from high-level ab initio calcns. on small mols. to semiempirical QM studies on a 1,2-dipalmitoyl-sn-phosphatidylcholine (DPPC) bilayer in combination with exptl. thermodn. data were used as target data for parameter optimization. These changes were tested with simulations of pure bilayers at high hydration of the following six lipids: DPPC, 1,2-dimyristoyl-sn-phosphatidylcholine (DMPC), 1,2-dilauroyl-sn-phosphatidylcholine (DLPC), 1-palmitoyl-2-oleoyl-sn-phosphatidylcholine (POPC), 1,2-dioleoyl-sn-phosphatidylcholine (DOPC), and 1-palmitoyl-2-oleoyl-sn-phosphatidylethanolamine (POPE); simulations of a low hydration DOPC bilayer were also performed. Agreement with exptl. surface area is on av. within 2%, and the d. profiles agree well with neutron and x-ray diffraction expts. NMR deuterium order parameters (SCD) are well predicted with the new FF, including proper splitting of the SCD for the aliph. carbon adjacent to the carbonyl for DPPC, POPE, and POPC bilayers. The area compressibility modulus and frequency dependence of 13C NMR relaxation rates of DPPC and the water distribution of low hydration DOPC bilayers also agree well with expt. Accordingly, the presented lipid FF, referred to as C36, allows for mol. dynamics simulations to be run in the tensionless ensemble (NPT), and is anticipated to be of utility for simulations of pure lipid systems as well as heterogeneous systems including membrane proteins. - 9Poger, D.; Van Gunsteren, W. F.; Mark, A. E. A new force field for simulating phosphatidylcholine bilayers J. Comput. Chem. 2010, 31 (6) 1117– 1125[Crossref], [PubMed], [CAS], Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjsFSkt7c%253D&md5=a1626ceb93b19b562b47a7c7a7d36b44A new force field for simulating phosphatidylcholine bilayersPoger, David; Van Gunsteren, Wilfred F.; Mark, Alan E.Journal of Computational Chemistry (2010), 31 (6), 1117-1125CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new force field for the simulation of dipalmitoylphosphatidylcholine (DPPC) in the liq.-cryst., fluid phase at zero surface tension is presented. The structure of the bilayer with the area per lipid (0.629 Nm2; expt. 0.629-0.64 Nm2), the vol. per lipid (1.226 Nm3; expt. 1.229-1.232 Nm3), and the ordering of the palmitoyl chains (order parameters) are all in very good agreement with expt. Exptl. electron d. profiles are well reproduced in particular with regard to the penetration of water into the bilayer. The force field was further validated by simulating the spontaneous assembly of DPPC into a bilayer in water. Notably, the timescale on which membrane sealing was obsd. using this model appears closer to the timescales for membrane resealing suggested by electroporation expts. than previous simulations using existing models. © 2009 Wiley Periodicals, Inc.
- 10Jämbeck, J. P. M.; Lyubartsev, A. P. Derivation and systematic validation of a refined all-atom force field for phosphatidylcholine lipids J. Phys. Chem. B 2012, 116 (10) 3164– 3179
- 11Marrink, S. J.; Risselada, H. J.; Yefimov, S.; Tieleman, D. P.; de Vries, A. H. The MARTINI force field: Coarse grained model for biomolecular simulations J. Phys. Chem. B 2007, 111 (27) 7812– 7824[ACS Full Text
], [CAS], Google Scholar
11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXmsVKmsLc%253D&md5=428d5750b94652e4917d905a30658235The MARTINI Force Field: Coarse Grained Model for Biomolecular SimulationsMarrink, Siewert J.; Risselada, H. Jelger; Yefimov, Serge; Tieleman, D. Peter; De Vries, Alex H.Journal of Physical Chemistry B (2007), 111 (27), 7812-7824CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present an improved and extended version of our coarse grained lipid model. The new version, coined the MARTINI force field, is parametrized in a systematic way, based on the reprodn. of partitioning free energies between polar and apolar phases of a large no. of chem. compds. To reproduce the free energies of these chem. building blocks, the no. of possible interaction levels of the coarse-grained sites has increased compared to those of the previous model. Application of the new model to lipid bilayers shows an improved behavior in terms of the stress profile across the bilayer and the tendency to form pores. An extension of the force field now also allows the simulation of planar (ring) compds., including sterols. Application to a bilayer/cholesterol system at various concns. shows the typical cholesterol condensation effect similar to that obsd. in all atom representations. - 12Orsi, M.; Essex, J. W. The ELBA force field for coarse-grain modeling of lipid membranes PLoS One 2011, 6 (12) e28637[Crossref], [PubMed], [CAS], Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjvFahsA%253D%253D&md5=feb2d1452c92ecce6a91dc2d28496690The ELBA force field for coarse-grain modeling of lipid membranesOrsi, Mario; Essex, Jonathan W.PLoS One (2011), 6 (12), e28637CODEN: POLNCL; ISSN:1932-6203. (Public Library of Science)A new coarse-grain model for mol. dynamics simulation of lipid membranes is presented. Following a simple and conventional approach, lipid mols. are modeled by spherical sites, each representing a group of several atoms. In contrast to common coarse-grain methods, two original (interdependent) features are here adopted. First, the main electrostatics are modeled explicitly by charges and dipoles, which interact realistically through a relative dielec. const. of unity (εr = 1). Second, water mols. are represented individually through a new parametrization of the simple Stockmayer potential for polar fluids; each water mol. is therefore described by a single spherical site embedded with a point dipole. The force field is shown to accurately reproduce the main phys. properties of single-species phospholipid bilayers comprising dioleoylphosphatidylcholine (DOPC) and dioleoylphosphatidylethanolamine (DOPE) in the liq. crystal phase, as well as distearoylphosphatidylcholine (DSPC) in the liq. crystal and gel phases. Insights are presented into fundamental properties and phenomena that can be difficult or impossible to study with alternative computational or exptl. methods. For example, we investigate the internal pressure distribution, dipole potential, lipid diffusion, and spontaneous self-assembly. Simulations lasting up to 1.5 μs were conducted for systems of different sizes (128, 512 and 1058 lipids); this also allowed us to identify size-dependent artifacts that are expected to affect membrane simulations in general. Future extensions and applications are discussed, particularly in relation to the methodol.'s inherent multiscale capabilities.
- 13Chiu, S.-W.; Pandit, S. A.; Scott, H. L.; Jakobsson, E. An improved united atom force field for simulation of mixed lipid bilayers J. Phys. Chem. B 2009, 113 (9) 2748– 2763[ACS Full Text
], [CAS], Google Scholar
13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Sltro%253D&md5=c881cacab451368c125262e4b9dd0e01An Improved United Atom Force Field for Simulation of Mixed Lipid BilayersChiu, See-Wing; Pandit, Sagar A.; Scott, H. L.; Jakobsson, EricJournal of Physical Chemistry B (2009), 113 (9), 2748-2763CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We introduce a new force field (43A1-S3) for simulation of membranes by the Gromacs simulation package. Construction of the force fields is by std. methods of electronic structure computations for bond parameters and charge distribution and sp. vol.s and heats of vaporization for small-mol. components of the larger lipid mols. for van der Waals parameters. Some parameters from the earlier 43A1 force field are found to be correct in the context of these calcns., while others are modified. The validity of the force fields is demonstrated by correct replication of X-ray form factors and NMR order parameters over a wide range of membrane compns. in semi-isotropic NTP 1 atm simulations. 43-A1-S3 compares favorably with other force fields used in conjunction with the Gromacs simulation package with respect to the breadth of phenomena that it accurately reproduces. - 14Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of multiple Amber force fields and development of improved protein backbone parameters Proteins: Struct., Funct., Bioinf. 2006, 65 (3) 712– 725
- 15Kirschner, K. N.; Yongye, A. B.; Tschampel, S. M.; González-Outeiriño, J.; Daniels, C. R.; Foley, B. L.; Woods, R. J. GLYCAM06: A generalizable biomolecular force field. Carbohydrates J. Comput. Chem. 2008, 29 (4) 622– 655
- 16Wang, 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, 1157– 1174[Crossref], [PubMed], [CAS], Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development 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.
- 17Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Swails, J.; Goetz, A. W.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Wolf, R. M.; Liu, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M.-J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Salomon-Ferrer, R.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A.AMBER 12; University of California: San Francisco, 2012.Google ScholarThere is no corresponding record for this reference.
- 18Salomon-Ferrer, R.; Case, D. A.; Walker, R. C. An overview of the Amber biomolecular simulation package Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2013, 3 (2) 198– 210[Crossref], [CAS], Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXmtFaiu70%253D&md5=baac68008e08e80bd825b2362e14fa1fAn overview of the amber biomolecular simulation packageSalomon-Ferrer, Romelia; Case, David A.; Walker, Ross C.Wiley Interdisciplinary Reviews: Computational Molecular Science (2013), 3 (2), 198-210CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. Mol. dynamics (MD) allows the study of biol. and chem. systems at the atomistic level on timescales from femtoseconds to milliseconds. It complements expt. while also offering a way to follow processes difficult to discern with exptl. techniques. Numerous software packages exist for conducting MD simulations of which one of the widest used is termed Amber. Here, we outline the most recent developments, since version 9 was released in Apr. 2006, of the Amber and AmberTools MD software packages, referred to here as simply the Amber package. The latest release represents six years of continued development, since version 9, by multiple research groups and the culmination of over 33 years of work beginning with the first version in 1979. The latest release of the Amber package, version 12 released in Apr. 2012, includes a substantial no. of important developments in both the scientific and computer science arenas. We present here a condensed vision of what Amber currently supports and where things are likely to head over the coming years. Figure 1 shows the performance in ns/day of the Amber package version 12 on a single-core AMD FX-8120 8-Core 3.6GHz CPU, the Cray XT5 system, and a single GPU GTX680.
- 19Götz, A. W.; Williamson, M. J.; Xu, D.; Poole, D.; Le Grand, S.; Walker, R. C. Routine microsecond molecular dynamics simulations with AMBER on GPUs. 1. Generalized Born J. Chem. Theory Comput. 2012, 8 (5) 1542– 1555[ACS Full Text
], [CAS], Google Scholar
19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XksFWns78%253D&md5=1e6f570db9cd504bb13706e7c56bc356Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 1. Generalized BornGotz, Andreas W.; Williamson, Mark J.; Xu, Dong; Poole, Duncan; Le Grand, Scott; Walker, Ross C.Journal of Chemical Theory and Computation (2012), 8 (5), 1542-1555CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present an implementation of generalized Born implicit solvent all-atom classical mol. dynamics (MD) within the AMBER program package that runs entirely on CUDA enabled NVIDIA graphics processing units (GPUs). We discuss the algorithms that are used to exploit the processing power of the GPUs and show the performance that can be achieved in comparison to simulations on conventional CPU clusters. The implementation supports three different precision models in which the contributions to the forces are calcd. in single precision floating point arithmetic but accumulated in double precision (SPDP), or everything is computed in single precision (SPSP) or double precision (DPDP). In addn. to performance, we have focused on understanding the implications of the different precision models on the outcome of implicit solvent MD simulations. We show results for a range of tests including the accuracy of single point force evaluations and energy conservation as well as structural properties pertaining to protein dynamics. The numerical noise due to rounding errors within the SPSP precision model is sufficiently large to lead to an accumulation of errors which can result in unphys. trajectories for long time scale simulations. We recommend the use of the mixed-precision SPDP model since the numerical results obtained are comparable with those of the full double precision DPDP model and the ref. double precision CPU implementation but at significantly reduced computational cost. Our implementation provides performance for GB simulations on a single desktop that is on par with, and in some cases exceeds, that of traditional supercomputers. - 20Salomon-Ferrer, R.; Götz, A. W.; Poole, D.; Le Grand, S.; Walker, R. C. Routine microsecond molecular dynamics simulations with Amber on GPUs. 2. Explicit solvent particle mesh Ewald J. Chem. Theory Comput. 2013, 9 (9) 3878– 3888[ACS Full Text
], [CAS], Google Scholar
20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1arsrzP&md5=57e18d35e8b81e9a79788ef5af17d19fRoutine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh EwaldSalomon-Ferrer, Romelia; Gotz, Andreas W.; Poole, Duncan; Le Grand, Scott; Walker, Ross C.Journal of Chemical Theory and Computation (2013), 9 (9), 3878-3888CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We present an implementation of explicit solvent all atom classical mol. dynamics (MD) within the AMBER program package that runs entirely on CUDA-enabled GPUs. First released publicly in Apr. 2010 as part of version 11 of the AMBER MD package and further improved and optimized over the last two years, this implementation supports the three most widely used statistical mech. ensembles (NVE, NVT, and NPT), uses particle mesh Ewald (PME) for the long-range electrostatics, and runs entirely on CUDA-enabled NVIDIA graphics processing units (GPUs), providing results that are statistically indistinguishable from the traditional CPU version of the software and with performance that exceeds that achievable by the CPU version of AMBER software running on all conventional CPU-based clusters and supercomputers. We briefly discuss three different precision models developed specifically for this work (SPDP, SPFP, and DPDP) and highlight the tech. details of the approach as it extends beyond previously reported work [Goetz et al., J. Chem. Theory Comput.2012, DOI: 10.1021/ct200909j; Le Grand et al., Comp. Phys. Comm.2013, DOI: 10.1016/j.cpc.2012.09.022].We highlight the substantial improvements in performance that are seen over traditional CPU-only machines and provide validation of our implementation and precision models. We also provide evidence supporting our decision to deprecate the previously described fully single precision (SPSP) model from the latest release of the AMBER software package. - 21Le Grand, S.; Götz, A. W.; Walker, R. C. SPFP: Speed without compromise—A mixed precision model for GPU accelerated molecular dynamics simulations Comput. Phys. Commun. 2013, 184 (2) 374– 380[Crossref], [CAS], Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsV2jsrzO&md5=2d9f102cb17f6e9bcbff3b5057d8c587SPFP: Speed without compromise-A mixed precision model for GPU accelerated molecular dynamics simulationsLe Grand, Scott; Gotz, Andreas W.; Walker, Ross C.Computer Physics Communications (2013), 184 (2), 374-380CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier B.V.)A new precision model is proposed for the acceleration of all-atom classical mol. dynamics (MD) simulations on graphics processing units (GPUs). This precision model replaces double precision arithmetic with fixed point integer arithmetic for the accumulation of force components as compared to a previously introduced model that uses mixed single/double precision arithmetic. This significantly boosts performance on modern GPU hardware without sacrificing numerical accuracy. We present an implementation for NVIDIA GPUs of both generalized Born implicit solvent simulations as well as explicit solvent simulations using the particle mesh Ewald (PME) algorithm for long-range electrostatics using this precision model. Tests demonstrate both the performance of this implementation as well as its numerical stability for const. energy and const. temp. biomol. MD as compared to a double precision CPU implementation and double and mixed single/double precision GPU implementations.
- 22Skjevik, Å. A.; Madej, B. D.; Walker, R. C.; Teigen, K. LIPID11: A modular framework for lipid simulations using Amber J. Phys. Chem. B 2012, 116 (36) 11124– 11136
- 23Siu, S. W.; Vacha, R.; Jungwirth, P.; Bockmann, R. A. Biomolecular simulations of membranes: physical properties from different force fields J. Chem. Phys. 2008, 128 (12) 125103[Crossref], [PubMed], [CAS], Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXktl2nsro%253D&md5=38ee4310c79b26dac45f24a14aa7d8e7Biomolecular simulations of membranes: Physical properties from different force fieldsSiu, Shirley W. I.; Vacha, Robert; Jungwirth, Pavel; Boeckmann, Rainer A.Journal of Chemical Physics (2008), 128 (12), 125103/1-125103/12CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Phospholipid force fields are of ample importance for the simulation of artificial bilayers, membranes, and also for the simulation of integral membrane proteins. Here, we compare the two most applied at. force fields for phospholipids, the all-atom CHARMM27 and the united atom Berger force field, with a newly developed all-atom generalized AMBER force field (GAFF) for dioleoylphosphatidylcholine mols. Only the latter displays the exptl. obsd. difference in the order of the C2 atom between the two acyl chains. The interfacial water dynamics is smoothly increased between the lipid carbonyl region and the bulk water phase for all force fields; however, the water order and with it the electrostatic potential across the bilayer showed distinct differences between the force fields. Both Berger and GAFF underestimate the lipid self-diffusion. GAFF offers a consistent force field for the at. scale simulation of biomembranes. (c) 2008 American Institute of Physics.
- 24Jójárt, B.; Martinek, T. A. Performance of the general amber force field in modeling aqueous POPC membrane bilayers J. Comput. Chem. 2007, 28 (12) 2051– 2058
- 25Rosso, L.; Gould, I. R. Structure and dynamics of phospholipid bilayers using recently developed general all-atom force fields J. Comput. Chem. 2008, 29 (1) 24– 37
- 26Dickson, C. J.; Rosso, L.; Betz, R. M.; Walker, R. C.; Gould, I. R. GAFFlipid: A general Amber force field for the accurate molecular dynamics simulation of phospholipid Soft Matter 2012, 8, 9617– 9627[Crossref], [CAS], Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1ymsbzL&md5=494ab83d6e79d084d2876c2169b9add1GAFFlipid: a General Amber Force Field for the accurate molecular dynamics simulation of phospholipidDickson, Callum J.; Rosso, Lula; Betz, Robin M.; Walker, Ross C.; Gould, Ian R.Soft Matter (2012), 8 (37), 9617-9627CODEN: SMOABF; ISSN:1744-683X. (Royal Society of Chemistry)Previous attempts to simulate phospholipid bilayers using the General Amber Force Field (GAFF) yielded many bilayer characteristics in agreement with expt., however when using a tensionless NPT ensemble the bilayer is seen to compress to an undesirable extent resulting in low areas per lipid and high order parameters in comparison to expt. In this work, the GAFF Lennard-Jones parameters for the simulation of acyl chains are cor. to allow the accurate and stable simulation of pure lipid bilayers. Lipid bilayers comprised of six phospholipid types were simulated for timescales approaching a quarter of a microsecond under tensionless const. pressure conditions using Graphics Processing Units. Structural properties including area per lipid, vol. per lipid, bilayer thickness, order parameter and headgroup hydration show favorable agreement with available exptl. values. Expanding the system size from 72 to 288 lipids and a more exptl. realistic 2 × 288 lipid bilayer stack induces little change in the obsd. properties. This preliminary work is intended for combination with the new AMBER Lipid11 modular force field as part of on-going attempts to create a modular phospholipid AMBER force field allowing tensionless NPT simulations of complex lipid bilayers.
- 27Siu, S. W. I.; Pluhackova, K.; Böckmann, R. A. Optimization of the OPLS-AA force field for long hydrocarbons J. Chem. Theory Comput. 2012, 8 (4) 1459– 1470[ACS Full Text
], [CAS], Google Scholar
27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjsFeiu78%253D&md5=aa2ce129ad72298695a988fe0e37af7dOptimization of the OPLS-AA Force Field for Long HydrocarbonsSiu, Shirley W. I.; Pluhackova, Kristyna; Boeckmann, Rainer A.Journal of Chemical Theory and Computation (2012), 8 (4), 1459-1470CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The all-atom optimized potentials for liq. simulations (OPLS-AA) force field is a popular force field for simulating biomols. However, the current OPLS parameters for hydrocarbons developed using short alkanes cannot reproduce the liq. properties of long alkanes in mol. dynamics simulations. Therefore, the extension of OPLS-AA to (phospho)lipid mols. required for the study of biol. membranes was hampered in the past. Here, we optimized the OPLS-AA force field for both short and long hydrocarbons. Following the framework of the OPLS-AA parametrization, we refined the torsional parameters for hydrocarbons by fitting to the gas-phase ab initio energy profiles calcd. at the accurate MP2/aug-cc-pVTZ theory level. Addnl., the depth of the Lennard-Jones potential for methylene hydrogen atoms was adjusted to reproduce the densities and the heats of vaporization of alkanes and alkenes of different lengths. Optimization of partial charges finally allowed to reproduce the gel-to-liq.-phase transition temp. for pentadecane and solvation free energies. The optimized parameter set (L-OPLS) yields improved hydrocarbon diffusion coeffs., viscosities, and gauche-trans ratios. Moreover, its applicability for lipid bilayer simulations is shown for a GMO bilayer in its liq.-cryst. phase. - 28Klauda, J. B.; Brooks, B. R.; MacKerell, A. D.; Venable, R. M.; Pastor, R. W. An ab initio study on the torsional surface of alkanes and its effect on molecular simulations of alkanes and a DPPC bilayer J. Phys. Chem. B 2005, 109 (11) 5300– 5311
- 29Betz, R. M.; Walker, R. C. Paramfit: Optimization of potential energy function parameters for molecular dynamics. Manuscript in preparation.Google ScholarThere is no corresponding record for this reference.
- 30Bayly, C. I.; Cieplak, P.; Cornell, W.; Kollman, P. A. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model J. Phys. Chem. 1993, 97 (40) 10269– 10280[ACS Full Text
], [CAS], Google Scholar
30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXlvVyqsLs%253D&md5=e65c6a556ffc174df4f327687912a0bdA well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP modelBayly, Christopher I.; Cieplak, Piotr; Cornell, Wendy; Kollman, Peter A.Journal of Physical Chemistry (1993), 97 (40), 10269-80CODEN: JPCHAX; ISSN:0022-3654.The authors present a new approach to generating electrostatic potential (ESP) derived charges for mols. The major strength of electrostatic potential derived charges is that they optimally reproduce the intermol. interaction properties of mols. with a simple two-body additive potential, provided, of course, that a suitably accurate level of quantum mech. calcn. is used to derive the ESP around the mol. Previously, the major weaknesses of these charges have been that they were not easily transferably between common functional groups in related mols., they have often been conformationally dependent, and the large charges that frequently occur can be problematic for simulating intramol. interactions. Introducing restraints in the form of a penalty function into the fitting process considerably reduces the above problems, with only a minor decrease in the quality of the fit to the quantum mech. ESP. Several other refinements in addn. to the restrained electrostatic potential (RESP) fit yield a general and algorithmic charge fitting procedure for generating atom-centered point charges. This approach can thus be recommended for general use in mol. mechanics, mol. dynamics, and free energy calcns. for any org. or bioorg. system. - 31Sonne, J.; Jensen, M. Ø.; Hansen, F. Y.; Hemmingsen, L.; Peters, G. H. Reparameterization of all-atom dipalmitoylphosphatidylcholine lipid parameters enables simulation of fluid bilayers at zero tension Biophys. J. 2007, 92 (12) 4157– 4167
- 32Klauda, J. B.; Garrison, S. L.; Jiang, J.; Arora, G.; Sandler, S. I. HM-IE: Quantum chemical hybrid methods for calculating interaction energies J. Phys. Chem. A 2003, 108 (1) 107– 112
- 33Davis, J. E.; Warren, G. L.; Patel, S. Revised charge equilibration potential for liquid alkanes J. Phys. Chem. B 2008, 112 (28) 8298– 8310
- 34Wang, J.; Hou, T. Application of molecular dynamics simulations in molecular property prediction. 1. density and heat of vaporization J. Chem. Theory Comput. 2011, 7 (7) 2151– 2165[ACS Full Text
], [CAS], Google Scholar
34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnt1Gmtbw%253D&md5=3485eec18e247b268e9733d189465f02Application of Molecular Dynamics Simulations in Molecular Property Prediction. 1. Density and Heat of VaporizationWang, Junmei; Hou, TingjunJournal of Chemical Theory and Computation (2011), 7 (7), 2151-2165CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Mol. mech. force field (FF) methods are useful in studying condensed phase properties. They are complementary to expts. and can often go beyond expts. in at. details. Even if a FF is specific for studying structures, dynamics, and functions of biomols., it is still important for the FF to accurately reproduce the exptl. liq. properties of small mols. that represent the chem. moieties of biomols. Otherwise, the force field may not describe the structures and energies of macromols. in aq. solns. properly. In this work, the authors have carried out a systematic study to evaluate the General AMBER Force Field (GAFF) in studying densities and heats of vaporization for a large set of org. mols. that covers the most common chem. functional groups. The latest techniques, such as the particle mesh Ewald (PME) for calcg. electrostatic energies and Langevin dynamics for scaling temps., have been applied in the mol. dynamics (MD) simulations. For d., the av. percent error (APE) of 71 org. compds. is 4.43% when compared with the exptl. values. More encouragingly, the APE drops to 3.43% after the exclusion of two outliers and four other compds. for which the exptl. densities have been measured with pressures higher than 1.0 atm. For the heat of vaporization, several protocols have been investigated, and the best one, P4/ntt0, achieves an av. unsigned error (AUE) and a root-mean-square error (RMSE) of 0.93 and 1.20 kcal/mol, resp. How to reduce the prediction errors through proper van der Waals (vdW) parametrization has been discussed. An encouraging finding in vdW parametrization is that both densities and heats of vaporization approach their "ideal" values in a synchronous fashion when vdW parameters are tuned. The following hydration free energy calcn. using thermodn. integration further justifies the vdW refinement. The authors conclude that simple vdW parametrization can significantly reduce the prediction errors. The authors believe that GAFF can greatly improve its performance in predicting liq. properties of org. mols. after a systematic vdW parametrization, which will be reported in a sep. paper. - 35Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N-log(N) method for Ewald sums in large systems J. Chem. Phys. 1993, 98 (12) 10089– 10092
- 36Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. C. Numerical integration of the cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes J. Comput. Phys. 1977, 23 (3) 327– 341[Crossref], [CAS], Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2sXktVGhsL4%253D&md5=b4aecddfde149117813a5ea4f5353ce2Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanesRyckaert, Jean Paul; Ciccotti, Giovanni; Berendsen, Herman J. C.Journal of Computational Physics (1977), 23 (3), 327-41CODEN: JCTPAH; ISSN:0021-9991.A numerical algorithm integrating the 3N Cartesian equation of motion of a system of N points subject to holonomic constraints is applied to mol. dynamics simulation of a liq. of 64 butane mols.
- 37Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R. Molecular dynamics with coupling to an external bath J. Chem. Phys. 1984, 81 (8) 3684– 3690[Crossref], [CAS], Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXmtlGksbY%253D&md5=5510dc00297d63b91ee3a7a4ae5aacb1Molecular dynamics with coupling to an external bathBerendsen, H. J. C.; Postma, J. P. M.; Van Gunsteren, W. F.; DiNola, A.; Haak, J. R.Journal of Chemical Physics (1984), 81 (8), 3684-90CODEN: JCPSA6; ISSN:0021-9606.In mol. dynamics (MD) simulations, the need often arises to maintain such parameters as temp. or pressure rather than energy and vol., or to impose gradients for studying transport properties in nonequil. MD. A method is described to realize coupling to an external bath with const. temp. or pressure with adjustable time consts. for the coupling. The method is easily extendable to other variables and to gradients, and can be applied also to polyat. mols. involving internal constraints. The influence of coupling time consts. on dynamical variables is evaluated. A leap-frog algorithm is presented for the general case involving constraints with coupling to both a const. temp. and a const. pressure bath.
- 38Yeh, I.-C.; Hummer, G. System-size dependence of diffusion coefficients and viscosities from molecular dynamics simulations with periodic boundary conditions J. Phys. Chem. B 2004, 108 (40) 15873– 15879[ACS Full Text
], [CAS], Google Scholar
38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXnsV2iurk%253D&md5=e333a4ea27b4451db997737f0e3f6a4fSystem-Size Dependence of Diffusion Coefficients and Viscosities from Molecular Dynamics Simulations with Periodic Boundary ConditionsYeh, In-Chul; Hummer, GerhardJournal of Physical Chemistry B (2004), 108 (40), 15873-15879CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We study the system-size dependence of translational diffusion coeffs. and viscosities in mol. dynamics simulations under periodic boundary conditions. Simulations of water under ambient conditions and a Lennard-Jones (LJ) fluid show that the diffusion coeffs. increase strongly as the system size increases. We test a simple analytic correction for the system-size effects that is based on hydrodynamic arguments. This correction scales as N-1/3, where N is the no. of particles. For a cubic simulation box of length L, the diffusion coeff. cor. for system-size effects is D0 = DPBC + 2.837297kBT/(6πηL), where DPBC is the diffusion coeff. calcd. in the simulation, kB the Boltzmann const., T the abs. temp., and η the shear viscosity of the solvent. For water, LJ fluids, and hard-sphere fluids, this correction quant. accounts for the system-size dependence of the calcd. self-diffusion coeffs. In contrast to diffusion coeffs., the shear viscosities of water and the LJ fluid show no significant system-size dependences. - 39Lipari, G.; Szabo, A. Effect of librational motion on fluorescence depolarization and nuclear magnetic resonance relaxation in macromolecules and membranes Biophys. J. 1980, 30 (3) 489– 506[Crossref], [PubMed], [CAS], Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXltV2gtLo%253D&md5=066330f119aa1ccabb117d13f54b02c4Effect of librational motion on fluorescence depolarization and nuclear magnetic resonance relaxation in macromolecules and membranesLipari, Giovanni; Szabo, AttilaBiophysical Journal (1980), 30 (3), 489-506CODEN: BIOJAU; ISSN:0006-3495.The theory of fluorescent emission anisotropy (r(t)) of a cylindrical probe in a membrane suspension is developed. The limiting anisotropy (r(∞)) is proportional to the square of the order parameter of the probe. The order parameter dets. the 1st nontrivial term in the expansion of the equil. orientational distribution function of the probe in a series of Legendre polynomials. The motion of the probe is described as diffusion (wobbling) within a cone of semiangle θ0. Within the framework of this model, an accurate single-exponential approxn. for r(t) is considered. An analytic expression relating the effective relaxation time, which appears in the above approxn., to θ0 and the diffusion coeff. for wobbling is derived. The model is generalized to the situation where the probe is attached to a macromol. whose motion cannot be neglected on the time scale of the fluorescence expt. Finally, by exploiting the formal similarity between the theory of fluorescence depolarization and 13C NMR dipolar relaxation, expressions for spin-spin and spin-lattice relaxation times and the nuclear Overhauser enhancement are derived for a protonated C which is nonrigidly attached to a macromol. and undergoes librational motion described as diffusion on a spherical cap of semiangle θ0.
- 40Ottiger, M.; Bax, A. Determination of Relative N–HN, N–C′, Cα–C′, and Cα–Hα effective bond lengths in a protein by NMR in a dilute liquid crystalline phase J. Am. Chem. Soc. 1998, 120 (47) 12334– 12341
- 41Haynes, W. M. CRC Handbook of Chemistry and Physics, 93rd ed.; CRC Press: Boca Raton, FL, 2012–2013.Google ScholarThere is no corresponding record for this reference.
- 42Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.Gaussian 09, Revision A.1; Gaussian Inc.: Wallingford, CT, 2009.Google ScholarThere is no corresponding record for this reference.
- 43Douglass, D. C.; McCall, D. W. Diffusion in paraffin hydrocarbons J. Phys. Chem. 1958, 62 (9) 1102– 1107
- 44Yaws, C. L. Yaws’ Handbook of Physical Properties for Hydrocarbons and Chemicals. http://www.knovel.com/web/portal/browse/display?_EXT_KNOVEL_DISPLAY_bookid=2147 (accessed February 19, 2013) .Google ScholarThere is no corresponding record for this reference.
- 45Tofts, P. S.; Lloyd, D.; Clark, C. A.; Barker, G. J.; Parker, G. J. M.; McConville, P.; Baldock, C.; Pope, J. M. Test liquids for quantitative MRI measurements of self-diffusion coefficient in vivo Magn. Reson. Med. 2000, 43 (3) 368– 374[Crossref], [PubMed], [CAS], Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXitVCgsbc%253D&md5=742855286475e83c3667d5a0bea824baTest liquids for quantitative MRI measurements of self-diffusion coefficient in vivoTofts, P. S.; Lloyd, D.; Clark, C. A.; Barker, G. J.; Parker, G. J. M.; McConville, P.; Baldock, C.; Pope, J. M.Magnetic Resonance in Medicine (2000), 43 (3), 368-374CODEN: MRMEEN; ISSN:0740-3194. (Wiley-Liss, Inc.)A range of liqs. suitable as quality control test objects for measuring the accuracy of clin. MRI diffusion sequences (both apparent diffusion coeff. and tensor) has been identified and characterized. The self-diffusion coeffs. for 15 liqs. (3 cyclic alkanes: cyclohexane to cyclooctane, 9 n-alkanes: n-octane to n-hexadecane, and 3 n-alcs.: ethanol to 1-propanol) were measured at 15-30° using an NMR spectrometer. Values at 22° range from 0.36 to 2.2 10-9 m2s-1. Typical 95% confidence limits are ±2%. Temp. coeffs. are 1.7-3.2%/°C. T1 and T2 values at 1.5 T and proton d. are given. N-tridecane has a diffusion coeff. close to that of normal white matter. The longer n-alkanes may be useful T2 stds. Measurements from a spin-echo MRI sequence agreed to within 2%.
- 46Holler, F.; Callis, J. B. Conformation of the hydrocarbon chains of sodium dodecyl sulfate molecules in micelles: an FTIR study J. Phys. Chem. 1989, 93 (5) 2053– 2058
- 47Lyerla, J. R.; McIntyre, H. M.; Torchia, D. A. A 13C nuclear magnetic resonance study of alkane motion Macromolecules 1974, 7 (1) 11– 14
- 48Jo, S.; Lim, J. B.; Klauda, J. B.; Im, W. CHARMM-GUI membrane builder for mixed bilayers and its application to yeast membranes Biophys. J. 2009, 97 (1) 50– 58[Crossref], [PubMed], [CAS], Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXpsFOmt7g%253D&md5=a6d57a8f42772e4c3674e324ea27ecaaCHARMM-GUI Membrane Builder for mixed bilayers and its application to yeast membranesJo, Sunhwan; Lim, Joseph B.; Klauda, Jeffery B.; Im, WonpilBiophysical Journal (2009), 97 (1), 50-58CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)The CHARMM-GUI Membrane Builder (http://www.charmm-gui.org/input/membrane), an intuitive, straightforward, web-based graphical user interface, was expanded to automate the building process of heterogeneous lipid bilayers, with or without a protein and with support for up to 32 different lipid types. The efficacy of these new features was tested by building and simulating lipid bilayers that resemble yeast membranes, composed of cholesterol, dipalmitoylphosphatidylcholine, dioleoylphosphatidylcholine, palmitoyloleoylphosphatidylethanolamine, palmitoyloleoylphosphatidylamine, and palmitoyloleoylphosphatidylserine. Four membranes with varying concns. of cholesterol and phospholipids were simulated, for a total of 170 ns at 303.15 K. Unsatd. phospholipid chain concn. had the largest influence on membrane properties, such as av. lipid surface area, d. profiles, deuterium order parameters, and cholesterol tilt angle. Simulations with a high concn. of unsatd. chains (73%, membraneunsat) resulted in a significant increase in lipid surface area and a decrease in deuterium order parameters, compared with membranes with a high concn. of satd. chains (60-63%, membranesat). The av. tilt angle of cholesterol with respect to bilayer normal was largest, and the distribution was significantly broader for membraneunsat. Moreover, short-lived cholesterol orientations parallel to the membrane surface existed only for membraneunsat. The membranesat simulations were in a liq.-ordered state, and agree with similar exptl. cholesterol-contg. membranes.
- 49Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water J. Chem. Phys. 1983, 79 (2) 926– 935[Crossref], [CAS], Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 50Joung, I. S.; Cheatham, T. E. Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations J. Phys. Chem. B 2008, 112 (30) 9020– 9041[ACS Full Text
], [CAS], Google Scholar
50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXnvFGqtL4%253D&md5=aa489470ae1c7479bf0911710217bd28Determination of Alkali and Halide Monovalent Ion Parameters for Use in Explicitly Solvated Biomolecular SimulationsJoung, In Suk; Cheatham, Thomas E.Journal of Physical Chemistry B (2008), 112 (30), 9020-9041CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Alkali (Li+, Na+, K+, Rb+, and Cs+) and halide (F-, Cl-, Br-, and I-) ions play an important role in many biol. phenomena, roles that range from stabilization of biomol. structure, to influence on biomol. dynamics, to key physiol. influence on homeostasis and signaling. To properly model ionic interaction and stability in atomistic simulations of biomol. structure, dynamics, folding, catalysis, and function, an accurate model or representation of the monovalent ions is critically necessary. A good model needs to simultaneously reproduce many properties of ions, including their structure, dynamics, solvation, and moreover both the interactions of these ions with each other in the crystal and in soln. and the interactions of ions with other mols. At present, the best force fields for biomols. employ a simple additive, nonpolarizable, and pairwise potential for at. interaction. In this work, the authors describe their efforts to build better models of the monovalent ions within the pairwise Coulombic and 6-12 Lennard-Jones framework, where the models are tuned to balance crystal and soln. properties in Ewald simulations with specific choices of well-known water models. Although it has been clearly demonstrated that truly accurate treatments of ions will require inclusion of nonadditivity and polarizability (particularly with the anions) and ultimately even a quantum mech. treatment, the authors' goal was to simply push the limits of the additive treatments to see if a balanced model could be created. The applied methodol. is general and can be extended to other ions and to polarizable force-field models. The authors' starting point centered on observations from long simulations of biomols. in salt soln. with the AMBER force fields where salt crystals formed well below their soly. limit. The likely cause of the artifact in the AMBER parameters relates to the naive mixing of the Smith and Dang chloride parameters with AMBER-adapted Aqvist cation parameters. To provide a more appropriate balance, the authors reoptimized the parameters of the Lennard-Jones potential for the ions and specific choices of water models. To validate and optimize the parameters, the authors calcd. hydration free energies of the solvated ions and also lattice energies (LE) and lattice consts. (LC) of alkali halide salt crystals. This is the first effort that systematically scans across the Lennard-Jones space (well depth and radius) while balancing ion properties like LE and LC across all pair combinations of the alkali ions and halide ions. The optimization across the entire monovalent series avoids systematic deviations. The ion parameters developed, optimized, and characterized were targeted for use with some of the most commonly used rigid and nonpolarizable water models, specifically TIP3P, TIP4PEW, and SPC/E. In addn. to well reproducing the soln. and crystal properties, the new ion parameters well reproduce binding energies of the ions to water and the radii of the first hydration shells. - 51Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. Numerical Recipes: The Art of Scientific Computing, 3rd ed. ed.; Cambridge University Press: New York, 2007.Google ScholarThere is no corresponding record for this reference.
- 52Pastor, R.; Brooks, B.; Szabo, A. An analysis of the accuracy of Langevin and molecular dynamics algorithms Mol. Phys. 1988, 65 (6) 1409– 1419
- 53Roe, D. R.; Cheatham, T. E. PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data J. Chem. Theory Comput. 2013, 9 (7) 3084– 3095[ACS Full Text
], [CAS], Google Scholar
53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXptFehtr8%253D&md5=6f1bee934f13f180bd7e1feb6b78036dPTRAJ and CPPTRAJ: Software for Processing and Analysis of Molecular Dynamics Trajectory DataRoe, Daniel R.; Cheatham, Thomas E.Journal of Chemical Theory and Computation (2013), 9 (7), 3084-3095CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We describe PTRAJ and its successor CPPTRAJ, two complementary, portable, and freely available computer programs for the anal. and processing of time series of three-dimensional at. positions (i.e., coordinate trajectories) and the data therein derived. Common tools include the ability to manipulate the data to convert among trajectory formats, process groups of trajectories generated with ensemble methods (e.g., replica exchange mol. dynamics), image with periodic boundary conditions, create av. structures, strip subsets of the system, and perform calcns. such as RMS fitting, measuring distances, B-factors, radii of gyration, radial distribution functions, and time correlations, among other actions and analyses. Both the PTRAJ and CPPTRAJ programs and source code are freely available under the GNU General Public License version 3 and are currently distributed within the AmberTools 12 suite of support programs that make up part of the Amber package of computer programs (see http://ambermd.org). This overview describes the general design, features, and history of these two programs, as well as algorithmic improvements and new features available in CPPTRAJ. - 54Kučerka, N.; Liu, Y.; Chu, N.; Petrache, H. I.; Tristram-Nagle, S.; Nagle, J. F. Structure of fully hydrated fluid phase DMPC and DLPC lipid bilayers using X-ray scattering from oriented multilamellar arrays and from unilamellar vesicles Biophys. J. 2005, 88 (4) 2626– 2637
- 55Kučerka, N.; Nieh, M.-P.; Katsaras, J. Fluid phase lipid areas and bilayer thicknesses of commonly used phosphatidylcholines as a function of temperature Biochim. Biophys. Acta 2011, 1808 (11) 2761– 2771
- 56Nagle, J. F. Introductory lecture: Basic quantities in model biomembranes Faraday Discuss. 2013, 161 (0) 11– 29
- 57Braun, A. R.; Sachs, J. N.; Nagle, J. F. Comparing simulations of lipid bilayers to scattering data: The GROMOS 43A1-S3 force field J. Phys. Chem. B 2013, 117 (17) 5065– 5072[ACS Full Text
], [CAS], Google Scholar
57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXltl2gs74%253D&md5=7d26d77825a6b983455aed5dac8321a9Comparing Simulations of Lipid Bilayers to Scattering Data: The GROMOS 43A1-S3 Force FieldBraun, Anthony R.; Sachs, Jonathan N.; Nagle, John F.Journal of Physical Chemistry B (2013), 117 (17), 5065-5072CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Simulations of DOPC at T = 303 K were performed using the united atom force field 43A1-S3 at six fixed projected areas, AP = 62, 64, 66, 68, 70, and 72 Å2, as well as a tensionless simulation that produced an av. ANPT = 65.8 Å2. After a small undulation correction for the system size consisting of 288 lipids, results were compared to exptl. data. The best, and excellent, fit to neutron scattering data occurs at an interpolated AN = 66.6 Å2 and the best, but not as good, fit to the more extensive x-ray scattering data occurs at AX = 68.7 Å2. The distance ΔDB-H between the Gibbs dividing surface for water and the peak in the electron d. profile agrees with scattering expts. The calcd. area compressibility KA = 277±10 mN/m is in excellent agreement with the micromech. expt. The vol. per lipid VL is smaller than vol. expts. which suggests a workaround that raises all the areas by about 1.5%. Although AX ≠ AN ≠ ANPT, this force field obtains acceptable agreement with expt. for AL = 67.5 Å2 (68.5 Å2 in the workaround), which we suggest is a better DOPC result from 43A1-S3 simulations than its value from the tensionless NPT simulation. However, nonsimulation modeling obtains better simultaneous fits to both kinds of scattering data, which suggests that the force fields can still be improved. - 58Rawicz, W.; Olbrich, K. C.; McIntosh, T.; Needham, D.; Evans, E. Effect of chain length and unsaturation on elasticity of lipid bilayers Biophys. J. 2000, 79 (1) 328– 339[Crossref], [PubMed], [CAS], Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1Kksrs%253D&md5=5d0a8c5e727c7620b846aebd6cc97cf5Effect of chain length and unsaturation on elasticity of lipid bilayersRawicz, W.; Olbrich, K. C.; McIntosh, T.; Needham, D.; Evans, E.Biophysical Journal (2000), 79 (1), 328-339CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Micropipette pressurization of giant bilayer vesicles was used to measure both elastic bending kc and area stretch KA moduli of fluid-phase phosphatidylcholine (PC) membranes. Twelve diacyl PCs were chosen: eight with two 18 carbon chains and degrees of unsatn. from one double bond (C18:1/0, C18:0/1) to six double bonds per lipid (diC18:3), two with short satd. carbon chains (diC13:0, diC14:0), and two with long unsatd. carbon chains (diC20:4, diC22:1). Bending moduli were derived from measurements of apparent expansion in vesicle surface area under very low tensions (0.001-0.5 mN/m), which is dominated by smoothing of thermal bending undulations. Area stretch moduli were obtained from measurements of vesicle surface expansion under high tensions (>0.5 mN/m), which involve an increase in area per mol. and a small, but important, contribution from smoothing of residual thermal undulations. The direct stretch moduli varied little (< ± 10%) with either chain unsatn. or length about a mean of 243 mN/m. On the other hand, the bending moduli of satd./mono-unsatd. chain PCs increased progressively with chain length from 0.56 × 10-19 J for diC13:0 to 1.2 × 10-19 J for diC22:1. However, quite unexpectedly for longer chains, the bending moduli dropped precipitously to ∼0.4 × 10-19 J when two or more cis double bonds were present in a chain (C18:0/2, diC18:2, diC18:3, diC20:4). Given nearly const. area stretch moduli, the variations in bending rigidity with chain length and poly-unsatn. implied significant variations in thickness. To test this hypothesis, peak-to-peak headgroup thicknesses hpp of bilayers were obtained from x-ray diffraction of multibilayer arrays at controlled relative humidities. For satd./mono-unsatd. chain bilayers, the distances hpp increased smoothly from diC13:0 to diC22:1 as expected. Moreover, the distances and elastic properties correlated well with a polymer brush model of the bilayer that specifies that the elastic ratio (kc/KA)1/2 = (hpp - ho)/24, where ho ≈ 1 nm accounts for sepn. of the headgroup peaks from the deformable hydrocarbon region. However, the elastic ratios and thicknesses for diC18:2, diC18:3, and diC20:4 fell into a distinct group below the correlation, which showed that poly-cis unsatd. chain bilayers are thinner and more flexible than satd./mono-unsatd. chain bilayers.
- 59Petrache, H. I.; Tristram-Nagle, S.; Nagle, J. F. Fluid phase structure of EPC and DMPC bilayers Chem. Phys. Lipids 1998, 95 (1) 83– 94
- 60Klauda, J. B.; Kučerka, N.; Brooks, B. R.; Pastor, R. W.; Nagle, J. F. Simulation-based methods for interpreting X-ray data from lipid bilayers Biophys. J. 2006, 90 (8) 2796– 2807[Crossref], [PubMed], [CAS], Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjsFeqtbw%253D&md5=727cd5ebed91aa22a6c6d1bff8000dd6Simulation-based methods for interpreting X-ray data from lipid bilayersKlauda, Jeffery B.; Kucerka, Norbert; Brooks, Bernard R.; Pastor, Richard W.; Nagle, John F.Biophysical Journal (2006), 90 (8), 2796-2807CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)The fully hydrated liq. cryst. phase of the dimyristoylphosphatidylcholine lipid bilayer at 30° was simulated using mol. dynamics with the CHARMM potential for five surface areas per lipid (A) in the range 55-65 Å2 that brackets the previously detd. exptl. area 60.6 Å2. The results of these simulations are used to develop a new hybrid zero-baseline structural model, denoted H2, for the electron d. profile, p(z), for the purpose of interpreting x-ray diffraction data. H2 and also the older hybrid baseline model were tested by fitting to partial information from the simulation and various constraints, both of which correspond to those available exptl. The A, ρ(z), and F(q) obtained from the models agree with those calcd. directly from simulation at each of the five areas, thereby validating this use of the models. The new H2 was then applied to exptl. dimyristoylphosphatidylcholine data; it yields A = 60.6 ± 0.5 Å2, in agreement with the earlier est. obtained using the hybrid baseline model. The electron d. profiles also compare well, despite considerable differences in the functional forms of the two models. Overall, the simulated ρ(z) at A = 60.7 Å2 agrees well with expt., demonstrating the accuracy of the CHARMM lipid force field; small discrepancies indicate targets for improvements. Lastly, a simulation-based model-free approach for obtaining A is proposed. It is based on interpolating the area that minimizes the difference between the exptl. F(q) and simulated F(q) evaluated for a range of surface areas. This approach is independent of structural models and could be used to det. structural properties of bilayers with different lipids, cholesterol, and peptides.
- 61Kučerka, N.; Tristram-Nagle, S.; Nagle, J. F. Closer look at structure of fully hydrated fluid phase DPPC bilayers Biophys. J. 2006, 90 (11) L83– L85
- 62Kučerka, N.; Nagle, J. F.; Sachs, J. N.; Feller, S. E.; Pencer, J.; Jackson, A.; Katsaras, J. Lipid bilayer structure determined by the simultaneous analysis of neutron and X-ray scattering data Biophys. J. 2008, 95 (5) 2356– 2367
- 63Evans, E.; Rawicz, W.; Smith, B. A. Concluding remarks back to the future: mechanics and thermodynamics of lipid biomembranes Faraday Discuss. 2013, 161 (0) 591– 611
- 64Evans, E.Personal Communication - DOPC isothermal compressibility modulus from X-ray data at 293 K; 2014.Google ScholarThere is no corresponding record for this reference.
- 65Tristram-Nagle, S.; Petrache, H. I.; Nagle, J. F. Structure and interactions of fully hydrated dioleoylphosphatidylcholine bilayers Biophys. J. 1998, 75 (2) 917– 925
- 66Pan, J.; Tristram-Nagle, S.; Kučerka, N.; Nagle, J. F. Temperature dependence of structure, bending rigidity, and bilayer interactions of dioleoylphosphatidylcholine bilayers Biophys. J. 2008, 94 (1) 117– 124
- 67Liu, Y.; Nagle, J. F. Diffuse scattering provides material parameters and electron density profiles of biomembranes Phys. Rev. E 2004, 69 (4) 040901
- 68Kučerka, N.; Tristram-Nagle, S.; Nagle, J. F. Structure of fully hydrated fluid phase lipid bilayers with monounsaturated chains J. Membr. Biol. 2006, 208 (3) 193– 202[Crossref], [CAS], Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjvF2lsbY%253D&md5=734e76569cbe3cf00f2a51d2a9543b7eStructure of Fully Hydrated Fluid Phase Lipid Bilayers with Monounsaturated ChainsKucerka, Norbert; Tristram-Nagle, Stephanie; Nagle, John F.Journal of Membrane Biology (2006), 208 (3), 193-202CODEN: JMBBBO; ISSN:0022-2631. (Springer)Quant. structures are obtained at 30°C for the fully hydrated fluid phases of palmitoyloleoylphosphatidylcholine (POPC), with a double bond on the sn-2 hydrocarbon chain, and for dierucoylphosphatidylcholine (di22:1PC), with a double bond on each hydrocarbon chain. The form factors F(qz) for both lipids are obtained using a combination of 3 methods: (1) Volumetric measurements provide F(0), (2) x-ray scattering from extruded unilamellar vesicles provides |F(qz)| for low qz, (3) Diffuse x-ray scattering from oriented stacks of bilayers provides |F(qz)| for high qz. Also, data using method (2) are added to the authors' recent data for dioleoylphosphatidylcholine (DOPC) using methods (1) and (3); the new DOPC data agree very well with the recent data and with (4) the authors' older data obtained using a liq. crystallog. x-ray method. The authors used hybrid electron d. models to obtain structural results from these form factors. The result for area per lipid (A) for DOPC 72.4 ± 0.5 Å2 agrees well with the authors' earlier publications, and the authors find A = 69.3 ± 0.5 Å2 for di22:1PC and A = 68.3 ± 1.5 Å2 for POPC. The authors obtain the values for 5 different av. thicknesses: hydrophobic, steric, head-head, phosphate-phosphate and Luzzati. Comparison of the results for these 3 lipids and for the authors' recent dimyristoylphosphatidylcholine (DMPC) detn. provides quant. measures of the effect of unsatn. on bilayer structure. The authors' results suggest that lipids with one monounsatd. chain have quant. bilayer structures closer to lipids with 2 monounsatd. chains than to lipids with 2 completely satd. chains.
- 69Binder, H.; Gawrisch, K. Effect of unsaturated lipid chains on dimensions, molecular order and hydration of membranes J. Phys. Chem. B 2001, 105 (49) 12378– 12390
- 70Rand, R. P.; Parsegian, V. A. Hydration forces between phospholipid bilayers Biochim. Biophys. Acta 1989, 988 (3) 351– 376
- 71Rappolt, M.; Hickel, A.; Bringezu, F.; Lohner, K. Mechanism of the lamellar/inverse hexagonal phase transition examined by high resolution X-ray diffraction Biophys. J. 2003, 84 (5) 3111– 3122[Crossref], [PubMed], [CAS], Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjvFGjur4%253D&md5=6a145c30169a2a14b0e664a625fbcf00Mechanism of the lamellar/inverse hexagonal phase transition examined by high resolution X-ray diffractionRappolt, Michael; Hickel, Andrea; Bringezu, Frank; Lohner, KarlBiophysical Journal (2003), 84 (5), 3111-3122CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)For the first time the electron d. of the lamellar liq. cryst. as well as of the inverted hexagonal phase could be retrieved at the transition temp. A reliable decompn. of the d-spacings into hydrophobic and hydrophilic structure elements could be performed owing to the presence of a sufficient no. of reflections. While the hydrocarbon chain length, dC, in the lamellar phase with a value of 14.5 Å lies within the extreme limits of the estd. chain length of the inverse hexagonal phase 10 Å < dC < 16 Å, the changes in the hydrophilic region vary strongly. During the lamellar-to-inverse hexagonal phase transition the area per lipid mol. reduces by ∼25%, and the no. of water mols. per lipid increases from 14 to 18. On the basis of the anal. of the structural components of each phase, the interface between the coexisting mesophases between 66 and 84° has been examd. in detail, and a model for the formation of the first rods in the matrix of the lamellar phospholipid stack is discussed. Judging from the structural relations between the inverse hexagonal and the lamellar phase, we suggest a cooperative chain reaction of rod formation at the transition midpoint, which is mainly driven by minimizing the interstitial region.
- 72Nagle, J. F.; Tristram-Nagle, S. Lipid bilayer structure Curr. Opin. Struct. Biol. 2000, 10 (4) 474– 480[Crossref], [PubMed], [CAS], Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXlvFOmt7Y%253D&md5=9cf3b9383d43db17029625724053da1eLipid bilayer structureNagle, John F.; Tristram-Nagle, StephanieCurrent Opinion in Structural Biology (2000), 10 (4), 474-480CODEN: COSBEF; ISSN:0959-440X. (Elsevier Science Ltd.)A review with 51 refs. Fluctuations, inherent in flexible and biol. relevant lipid bilayers, make quant. structure detn. challenging. Shortcomings in older methods of structure detn. have been realized and new methodologies have been introduced that take fluctuations into account. The large uncertainty in literature values of lipid bilayer structural parameters is being reduced.
- 73Anézo, C.; de Vries, A. H.; Höltje, H.-D.; Tieleman, D. P.; Marrink, S.-J. Methodological issues in lipid bilayer simulations J. Phys. Chem. B 2003, 107 (35) 9424– 9433
- 74Poger, D.; Mark, A. E. On the validation of molecular dynamics simulations of saturated and cis-monounsaturated phosphatidylcholine lipid bilayers: A comparison with experiment J. Chem. Theory Comput. 2009, 6 (1) 325– 336
- 75Kučerka, N.; Katsaras, J.; Nagle, J. Comparing membrane simulations to scattering experiments: Introducing the SIMtoEXP software J. Membr. Biol. 2010, 235 (1) 43– 50[Crossref], [PubMed], [CAS], Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXls1Wkurg%253D&md5=52e94861b62db6b538f7dc502b680818Comparing Membrane Simulations to Scattering Experiments: Introducing the SIMtoEXP SoftwareKucerka, Norbert; Katsaras, John; Nagle, John F.Journal of Membrane Biology (2010), 235 (1), 43-50CODEN: JMBBBO; ISSN:0022-2631. (Springer)SIMtoEXP is a software package designed to facilitate the comparison of biomembrane simulations with exptl. x-ray and neutron scattering data. It has the following features: (1) Accepts no. d. profiles from simulations in a std. but flexible format. (2) Calcs. the electron d. ε(z) and neutron scattering length d. ν(z) profiles along the z direction (i.e., normal to the membrane) and their resp. Fourier transforms (i.e., Fε [qz] and Fν[qz]). The resultant 4 functions are then displayed graphically. (3) Accepts exptl. Fε(qz) and Fν(qz) data for graphical comparison with simulations. (4) Allows for lipids and other large mols. to be parsed into component groups by the user and calcs. the component vols. following Petrache et al. The software then calcs. and displays the contributions of each component group as vol. probability profiles, ρ(z), as well as the contributions of each component to ε(z) and ν(z).
- 76Shirts, M. R. Simple quantitative tests to validate sampling from thermodynamic ensembles J. Chem. Theory Comput. 2012, 9 (2) 909– 926
- 77Seelig, J.; Waespe-Sarcevic, N. Molecular order in cis and trans unsaturated phospholipid bilayers Biochemistry 1978, 17 (16) 3310– 3315
- 78Perly, B.; Smith, I. C. P.; Jarrell, H. C. Acyl chain dynamics of phosphatidylethanolamines containing oleic acid and dihydrosterculic acid: deuteron NMR relaxation studies Biochemistry 1985, 24 (17) 4659– 4665
- 79Lafleur, M.; Bloom, M.; Eikenberry, E. F.; Gruner, S. M.; Han, Y.; Cullis, P. R. Correlation between lipid plane curvature and lipid chain order Biophys. J. 1996, 70 (6) 2747– 2757[Crossref], [PubMed], [CAS], Google Scholar79https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVCrtb4%253D&md5=ff34b0b6df6c340bdc716572baf34935Correlation between lipid plane curvature and lipid chain orderLafleur, Michel; Bloom, Myer; Eikenberry, Eric F.; Gruner, Sol M.; Han, Yuqi; Cullis, Pieter R.Biophysical Journal (1996), 70 (6), 2747-2757CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)The 1-palmitoyl-2-oleoyl-phosphatidylethanolamine: 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPE:POPC) system has been investigated by measuring, in the inverted hexagonal (HII) phase, the intercylinder spacings (using x-ray diffraction) and orientational order of the acyl chains (using 2H NMR). The presence of 20% dodecane leads to the formation of a HII phase for the compn. range from 0 to 39 mol% of POPC in POPE, as ascertained by x-ray diffraction and 2H NMR. The addn. of the alkane induces a small decrease in chain order, consistent with less stretched chains. An increase in temp. or in POPE proportion leads to a redn. in the intercylinder spacing, primarily due to a decrease in the water core radius. A temp. increase also leads to a redn. in the orientational order of the lipid acyl chains, whereas in POPE proportion has little effect on chain order. A correlation is proposed to relate the radius of curvature of the cylinders in the inverted hexagonal phase to the chain order of the lipids adopting the HII phase. A simple geometrical model is proposed, taking into account the area occupied by the polar headgroup at the interface and the orientational order of the acyl chains reflecting the contribution of the apolar core. From these parameters, intercylinder spacings are calcd. that agree well with the values detd. exptl. by x-ray diffraction, for the variations of both temp. and POPE:POPC proportion. This model suggests that temp. increases the curvature of lipid layers, mainly by increasing the area subtended by the hydrophobic core through chain conformation disorder, whereas POPC content affects primarily the headgroup interface contribution. The frustration of lipid layer curvature is also shown to be reflected in the acyl chain order measured in the Lα phase, in the absence of dodecane; for a given temp., increased order is obsd. when the curling tendencies of the lipid plane are more pronounced.
- 80Warschawski, D.; Devaux, P. Order parameters of unsaturated phospholipids in membranes and the effect of cholesterol: a 1H-13C solid-state NMR study at natural abundance Eur. Biophys. J. 2005, 34 (8) 987– 96[Crossref], [PubMed], [CAS], Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SksbvJ&md5=b5ac0c99dd2dad942a2c48c20e8516c5Order parameters of unsaturated phospholipids in membranes and the effect of cholesterol: a 1H-13C solid-state NMR study at natural abundanceWarschawski, Dror E.; Devaux, Philippe F.European Biophysics Journal (2005), 34 (8), 987-996CODEN: EBJOE8; ISSN:0175-7571. (Springer)Most biol. phospholipids contain at least one unsatd. alkyl chain. However, few order parameters of unsatd. lipids have been detd. because of the difficulty assocd. with isotopic labeling of a double bond. Dipolar recoupling on axis with scaling and shape preservation (DROSS) is a solid-state NMR technique optimized for measuring 1H-13C dipolar couplings and order parameters in lipid membranes in the fluid phase. It has been used to det. the order profile of 1,2-dimyristoyl-sn-glycero-3-phosphocholine hydrated membranes. Here, we show an application for the measurement of local order parameters in multilamellar vesicles contg. unsatd. lipids. Taking advantage of the very good 13C chem. shift dispersion, one can easily follow the segmental order along the acyl chains and, particularly, around the double bonds where we have been able to det. the previously misassigned order parameters of each acyl chain of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC). We have followed the variation of such order profiles with temp., unsatn. content and cholesterol addn. We have found that the phase formed by DOPC with 30% cholesterol is analogous to the liq.-ordered (lo) phase. Because these expts. do not require isotopic enrichment, this technique can, in principle, be applied to natural lipids and biomembranes.
- 81Petrache, H. I.; Dodd, S. W.; Brown, M. F. Area per lipid and acyl length distributions in fluid phosphatidylcholines determined by (2)H NMR spectroscopy Biophys. J. 2000, 79 (6) 3172– 92[Crossref], [PubMed], [CAS], Google Scholar81https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXitFCn&md5=26c87135c39ab21617ea1987653b083fArea per lipid and acyl length distributions in fluid phosphatidylcholines determined by 2H NMR spectroscopyPetrache, Horia I.; Dodd, Steven W.; Brown, Michael F.Biophysical Journal (2000), 79 (6), 3172-3192CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)Deuterium (2H) NMR spectroscopy provides detailed information regarding the structural fluctuations of lipid bilayers, including both the equil. properties and dynamics. Exptl. 2H NMR measurements for the homologous series of 1,2-diacyl-sn-glycero-3-phosphocholines with perdeuterated satd. chains (from C12:0 to C18:0) have been performed on randomly oriented, fully hydrated multilamellar samples. For each lipid, the C-D bond order parameters have been calcd. from de-Paked 2H NMR spectra as a function of temp. The exptl. order parameters were analyzed using a mean-torque potential model for the acyl chain segment distributions, and comparison was made with the conventional diamond lattice approach. Statistical mech. principles were used to relate the measured order parameters to the lipid bilayer structural parameters: the hydrocarbon thickness and the mean interfacial area per lipid. At fixed temp., the area decreases with increasing acyl length, indicating increased van der Waals attraction for longer lipid chains. However, the main effect of increasing the acyl chain length is on the hydrocarbon thickness rather than on the area per lipid. Expansion coeffs. of the structural parameters are reported and interpreted using an empirical free energy function that describes the force balance in fluid bilayers. At the same abs. temp., the phosphatidylcholine (PC) series exhibits a universal chain packing profile that differs from that of phosphatidylethanolamines (PE). Hence, the lateral packing of phospholipids is more sensitive to the headgroup methylation than to the acyl chain length. A fit to the area per lipid for the PC series using the empirical free energy function shows that the PE area represents a limiting value for the packing of fluid acyl chains.
- 82Douliez, J. P.; Léonard, A.; Dufourc, E. J. Restatement of order parameters in biomembranes: calculation of C-C bond order parameters from C-D quadrupolar splittings Biophys. J. 1995, 68 (5) 1727– 1739[Crossref], [PubMed], [CAS], Google Scholar82https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXlt1KgsrY%253D&md5=cb2fa7e7b26000ab88e390ee0bd25090Restatement of order parameters in biomembranes: calculation of C-C bond order parameters from C-D quadrupolar splittingsDouliez, Jean-Paul; Leonard, Alain; Dufourc, Erick J.Biophysical Journal (1995), 68 (5), 1727-39CODEN: BIOJAU; ISSN:0006-3495. (Biophysical Society)An expression for the C-C bond order parameter, SCC, of membrane hydrocarbon chains has been derived from the obsd. C-D bond order parameters. It allows calcn. of the probability of each of the C-C bond rotamers and, consequently, the no. of gauche defects per chain as well as their projected av. length onto the bilayer normal, thus affording the calcn. of accurate hydrophobic bilayer thicknesses. The effect of the temp. has been studied on dilauroyl-, dimyristoyl-, and dipalmitoylphosphatidylcholine (DLPC, DMPC, DPPC) membranes, as has the effect of cholesterol on DMPC. The salient results are as follows: (1) an odd-even effect is obsd. for the SCC vs. carbon position, k, whose amplitude increases with temp.; (2) calcn. of SCC, from nonequivalent deuterons on the sn-2 chain of lipids, S2CC, leads to neg. values, indicating the tendency for the C1-C2 bond to be oriented parallel to the bilayer surface; this bond becomes more parallel to the surface as the temp. increases or when cholesterol is added; (3) calcn. on the sn-2 chain length can be performed from C1 and Cn, where n is the no. of carbon atoms in the chain, and leads to 10.4, 12.2, and 13.8 Å for DLPC, DMPC, and DPPC close to the transition temp. TC, of each of the systems and to 9.4, 10.9, and 12.6 for T-TC = 30-40°, resp.; (4) sepn. of intra- and intermol. motions allows quantitation of the no. of gauche defects per chain, which is equal to 1.9, 2.7, and 3.5 for DLPC, DMPC, and DPPC near Tc and to 2.7, 3.5, and 4.4 at T - TC = 30-40°, resp. Finally, the validity of the model is discussed and compared with previously published models.
- 83Aussenac, F.; Laguerre, M.; Schmitter, J.-M.; Dufourc, E. J. Detailed structure and dynamics of bicelle phospholipids using selectively deuterated and perdeuterated labels. 2H NMR and molecular mechanics study Langmuir 2003, 19 (25) 10468– 10479
- 84Shaikh, S. R.; Brzustowicz, M. R.; Gustafson, N.; Stillwell, W.; Wassall, S. R. Monounsaturated PE does not phase-separate from the lipid raft molecules sphingomyelin and cholesterol: Role for polyunsaturation? Biochemistry 2002, 41 (34) 10593– 10602
- 85Hitchcock, P. B.; Mason, R.; Thomas, K. M.; Shipley, G. G. Structural chemistry of 1,2 dilauroyl-DL-phosphatidylethanolamine: Molecular conformation and intermolecular packing of phospholipids Proc. Natl. Acad. Sci. U.S.A. 1974, 71 (8) 3036– 3040
- 86Seelig, A.; Seelig, J. Bilayers of dipalmitoyl-3-sn-phosphatidylcholine: Conformational differences between the fatty acyl chains Biochim. Biophys. Acta 1975, 406 (1) 1– 5
- 87Jämbeck, J. P. M.; Lyubartsev, A. P. An extension and further validation of an all-atomistic force field for biological membranes J. Chem. Theory Comput. 2012, 8 (8) 2938– 2948
- 88Senak, L.; Davies, M. A.; Mendelsohn, R. A quantitative IR study of hydrocarbon chain conformation in alkanes and phospholipids: CH2 wagging modes in disordered bilayer and HII phases J. Phys. Chem. 1991, 95 (6) 2565– 2571[ACS Full Text
], [CAS], Google Scholar
88https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhvFemsLw%253D&md5=86298b46c827e9371a74b49f9dd28aefA quantitative IR study of hydrocarbon chain conformation in alkanes and phospholipids: CH2 wagging modes in disordered bilayer and HII phasesSenak, Laurence; Davies, Mark A.; Mendelsohn, RichardJournal of Physical Chemistry (1991), 95 (6), 2565-71CODEN: JPCHAX; ISSN:0022-3654.A series of CH2 wagging modes in the 1320-1370-cm-1 region of alkane and phospholipid IR spectra characteristic of nonplanar conformers were used for quant. evaluation of conformational states in disordered (phospholipid Lα and HII and alkane liq.) phases. A quant. comparison of DPPC and 1,2-dipalmitoylphosphatidylethanolamine (DPPE) shows the former to have 0.4 double gauche (gg), 0.5 end gauche (eg), and ∼1.0 (kink + gtg) conformers per chain just above the gel0liq. crystal phase transition, while the more highly ordered DPPE Lα phase shows ∼0.2 gg, 0.1 eg, and 1.0 (kink + gtg) conformers. In all systems studied, the no. of allowed gg and eg forms that occur is reduced substantially. From those achieved in isotropic liq. alkanes. A study of the Lα → HII interconversion in 2 unsatd. (1-palmitoyl-2-oleoyl- and 1,2-dielaidoylphosphatidylethanolamine) shows a substantial increase in both gg and eg conformers near temp.3s leading to the inverted micellar state in each instance, with smaller percentage increases in (kink + gtg) states. Connections between these quant. observations with those of 2H NMR and other spectroscopies are presented. Finally, a temp.4 dependence inconsistent with the prediction of the rotational isomeric state model is noted for the 1368-cm-1 band characteristic of (kink + gtg) conformers in alkanes. - 89Moss, G. P. Basic terminology of stereochemistry Pure Appl. Chem. 1996, 68 (12) 2193– 2222
- 90Cates, D. A.; Strauss, H. L.; Snyder, R. G. Vibrational modes of liquid n-alkanes: Simulated isotropic raman spectra and band progressions for C5H12-C20H42 and C16D34 J. Phys. Chem. 1994, 98 (16) 4482– 4488
- 91Snyder, R. G.; Strauss, H. L.; Elliger, C. A. C-H stretching modes and the structure of n-alkyl chains. 1. Long, disordered chains J. Phys. Chem. 1982, 86, 5145– 5150
- 92Mendelsohn, R.; Senak, L. Quantitative determination of conformational disorder in biological membranes by FTIR spectroscopy. In Biomolecular spectroscopy; Clark, R. J. H.; Hester, R. E., Eds.; Wiley: New York, 1993; pp 339– 380.Google ScholarThere is no corresponding record for this reference.
- 93Casal, H. L.; McElhaney, R. N. Quantitative determination of hydrocarbon chain conformational order in bilayers of saturated phosphatidylcholines of various chain lengths by Fourier transform infrared spectroscopy Biochemistry 1990, 29 (23) 5423– 5427
- 94Tuchtenhagen, J.; Ziegler, W.; Blume, A. Acyl chain conformational ordering in liquid-crystalline bilayers: comparative FT-IR and 2H-NMR studies of phospholipids differing in headgroup structure and chain length Eur. Biophys. J. 1994, 23 (5) 323– 335[Crossref], [CAS], Google Scholar94https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXltFClurg%253D&md5=44dcd68c3e24ac87b38aacf1dc9a1338Acyl chain conformational ordering in liquid-crystalline bilayers: comparative FT-IR and 2H-NMR studies of phospholipids differing in headgroup structure and chain lengthTuchtenhagen, Juergen; Ziegler, Wolfgang; Blume, AlfredEuropean Biophysics Journal (1994), 23 (5), 323-35CODEN: EBJOE8; ISSN:0175-7571.FT-IR spectroscopy has been used to evaluate the acyl chain conformational ordering of DMPC, DMPE, DMPA (pH 6 and 12), DMPG (pH 1 and 7), and DPPC, DPPE, DPPA (pH 6). The frequencies of the sym. and antisym. methylene stretching vibrations were detd. as a function of temp. In the liq.-cryst. phase the frequencies show a qual. dependence on the amt. of chain disorder. Quant. data for trans-gauche isomerization were obtained from the integral intensities of the conformation sensitive methylene wagging absorptions at ∼1368 cm-1 (gtg' and gtg sequences), 1356 cm-1 (double gauche) and 1342 cm-1 (end gauche). The integral band intensities were converted to the no. of gauche conformers per acyl chain using the calibration factors published by L. Senak et al. (1991). At 69° the highest no. of gauche conformers excluding contributions from single gauche conformers and jogs (g'tttg) were found for PCs (DMPC: 2.6; DPPC: 2.4), followed by DMPG- (2.0), phosphatidylethanolamines (DMPE: 1.4; DPPE: 2.0), protonated DMPG (1.5), and phosphatidic acids (DPPA-: 1.7; DMPA-: 1.4, DMPA2-: 1.7).
- 95Mendelsohn, R.; Davies, M. A.; Brauner, J. W.; Schuster, H. F.; Dluhy, R. A. Quantitative determination of conformational disorder in the acyl chains of phospholipid bilayers by infrared spectroscopy Biochemistry 1989, 28 (22) 8934– 8939
- 96Kučerka, N.; Gallová, J.; Uhríková, D.; Balgavý, P.; Bulacu, M.; Marrink, S.-J.; Katsaras, J. Areas of Monounsaturated Diacylphosphatidylcholines Biophys. J. 2009, 97 (7) 1926– 1932
- 97Ratto, T. V.; Longo, M. L. Obstructed diffusion in phase-separated supported lipid bilayers: A combined atomic force microscopy and fluorescence recovery after photobleaching approach Biophys. J. 2002, 83 (6) 3380– 3392
- 98Almeida, P. F. F.; Vaz, W. L. C.; Thompson, T. E. Lateral diffusion in the liquid phases of dimyristoylphosphatidylcholine/cholesterol lipid bilayers: a free volume analysis Biochemistry 1992, 31 (29) 6739– 6747[ACS Full Text
], [CAS], Google Scholar
98https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38Xks1Gisbc%253D&md5=1478036ef44a66d02c70da8e2dd0ff8cLateral diffusion in the liquid phases of dimyristoylphosphatidylcholine/cholesterol lipid bilayers: a free volume analysisAlmeida, Paulo F. F.; Vaz, Winchil L. C.; Thompson, T. E.Biochemistry (1992), 31 (29), 6739-47CODEN: BICHAW; ISSN:0006-2960.Fluorescence recovery after photobleaching is used to perform an extensive study of the lateral diffusion of a phospholipid probe in the binary mixt. dimyristoylphosphatidylcholine/cholesterol, above the melting temp. of the phospholipid. In the regions of the phase diagram where a single liq. phase exists, diffusion can be quant. described by free vol. theory, using a modified Macedo-Litovitz hybrid equation. In the liq.-liq. immiscibility region, the temp. dependence of the diffusion coeff. is in excellent agreement with current theories of generalized diffusivities in composite two-phase media. A consistent interpretation of the diffusion data can be provided based essentially on the idea that the primary effect of cholesterol addn. to the bilayer is to occupy free vol. On this basis, a general interpretation of the phase behavior of this mixt. is also proposed. - 99Orädd, G.; Lindblom, G.; Westerman, P. W. Lateral diffusion of cholesterol and dimyristoylphosphatidylcholine in a lipid bilayer measured by pulsed field gradient NMR spectroscopy Biophys. J. 2002, 83 (5) 2702– 2704
- 100Filippov, A.; Orädd, G.; Lindblom, G. Influence of cholesterol and water content on phospholipid lateral diffusion in bilayers Langmuir 2003, 19 (16) 6397– 6400
- 101Vaz, W. L. C.; Clegg, R. M.; Hallmann, D. Translational diffusion of lipids in liquid crystalline phase phosphatidylcholine multibilayers. A comparison of experiment with theory Biochemistry 1985, 24 (3) 781– 786[ACS Full Text
], [CAS], Google Scholar
101https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXptFKqsA%253D%253D&md5=5c01282a25df5150530f13a60cbc9dddTranslational diffusion of lipids in liquid crystalline phase phosphatidylcholine multibilayers. A comparison of experiment with theoryVaz, Winchil L. C.; Clegg, Robert M.; Hallmann, DieterBiochemistry (1985), 24 (3), 781-6CODEN: BICHAW; ISSN:0006-2960.A systematic study of the translational diffusion of the phospholipid deriv., N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine (NBD-PE), was undertaken in liq. cryst. phase phosphatidylcholine (PC) bilayers by using the fluorescence recovery after photobleaching technique. The exptl. results were then compared with the predictions of theor. models for diffusion in membranes. For NBD-PE, the dependence of the translational diffusion coeff. (Dt) upon the acyl chain length of the diffusant was not that predicted by continuum fluid hydrodynamic models for diffusion in membranes. Plots of Dt vs. 1/T (Arrhenius plots) were nonlinear in PC bilayers where the acyl chain compn. of the NBD-PE was matched with that of the host bilayer lipid. This suggested that a free vol. model may be appropriate for the description of lipid diffusion in lipid bilayers. In bilayers of PCs with satd. acyl chains at the same reduced temp., the magnitude of Dt followed the order: distearoyl-PC > dipalmitoyl-PC (DPPC) > dimyristoyl-PC (DMPC) > dilauroyl-PC (DLPC). This was the inverse of what might be expected from the hydrodynamic model, but was in agreement with the free vol. in these bilayers. A free vol. model that takes into account the frictional drag forces acting upon the diffusing NBD-PE at the membrane-water interface and also at the bilayer midplane adequately describes the diffusion results for NBD-PE mols. in DLPC, DMPC, DPPC, and 1-palmitoyl-2-oleoyl-PC bilayers in the liq.-cryst. phase. - 102Scheidt, H. A.; Huster, D.; Gawrisch, K. Diffusion of cholesterol and its precursors in lipid membranes studied by 1H pulsed field gradient magic angle spinning NMR Biophys. J. 2005, 89 (4) 2504– 2512
- 103Kuśba, J.; Li, L.; Gryczynski, I.; Piszczek, G.; Johnson, M.; Lakowicz, J. R. Lateral diffusion coefficients in membranes measured by resonance energy transfer and a new algorithm for diffusion in two dimensions Biophys. J. 2002, 82 (3) 1358– 1372
- 104Jin, A. J.; Edidin, M.; Nossal, R.; Gershfeld, N. L. A singular state of membrane lipids at cell growth temperatures Biochemistry 1999, 38 (40) 13275– 13278[ACS Full Text
], [CAS], Google Scholar
104https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXlvVSisrs%253D&md5=ae149f4a53ff9c3f3e906f6763212046A singular state of membrane lipids at cell growth temperaturesJin, A. J.; Edidin, M.; Nossal, R.; Gershfeld, N. L.Biochemistry (1999), 38 (40), 13275-13278CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Cells adjust their membrane lipid compn. when they adapt to grow at different temps. The consequences of this adjustment for membrane properties and functions are not well understood. Our report shows that the temp. dependence of the diffusion of a probe mol. in multilayers formed from total lipid exts. of E. coli has an anomalous max. at a temp. corresponding to the growth temp. of each bacterial prepn. (25, 29, and 32 °C). This increase in the lateral diffusion coeff., D, is characteristic of membrane lipids in a crit. state, for which large fluctuations of mol. area in the plane of the bilayer are expected. Therefore, changes in lipid compn. may be due to a requirement that cells maintain their membranes in a state where mol. interactions and reaction rates are readily modulated by small, local perturbations of membrane organization. - 105Poger, D.; Mark, A. E. Lipid bilayers: The effect of force field on ordering and dynamics J. Chem. Theory Comput. 2012, 8 (11) 4807– 4817[ACS Full Text
], [CAS], Google Scholar
105https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVOhs7vI&md5=2870d3c59c4b8baa96dae55ca46b3a75Lipid Bilayers: The Effect of Force Field on Ordering and DynamicsPoger, David; Mark, Alan E.Journal of Chemical Theory and Computation (2012), 8 (11), 4807-4817CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The sensitivity of the structure and dynamics of a fully hydrated pure bilayer of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) in mol. dynamics simulations to changes in force-field and simulation parameters has been assessed. Three related force fields (the GROMOS 54A7 force field, a GROMOS 53A6-derived parameter set and a variant of the Berger parameters) in combination with either particle-mesh Ewald (PME) or a reaction field (RF) were compared. Structural properties such as the area per lipid, carbon-deuterium order parameters, electron d. profile and bilayer thicknesses, are reproduced by all the parameter sets within the uncertainty of the available exptl. data. However, there are clear differences in the ordering of the glycerol backbone and choline headgroup, and the orientation of the headgroup dipole. In some cases, the degree of ordering was reminiscent of a liq.-ordered phase. It is also shown that, although the lateral diffusion of the lipids in the plane of the bilayer is often used to validate lipid force fields, because of the uncertainty in the exptl. measurements and the fact that the lateral diffusion is dependent on the choice of the simulation conditions, it should not be employed as a measure of quality. Finally, the simulations show that the effect of small changes in force-field parameters on the structure and dynamics of a bilayer is more significant than the treatment of the long-range electrostatic interactions using RF or PME. Overall, the GROMOS 54A7 best reproduced the range of exptl. data examd. - 106Wohlert, J.; Edholm, O. Dynamics in atomistic simulations of phospholipid membranes: Nuclear magnetic resonance relaxation rates and lateral diffusion J. Chem. Phys. 2006, 125 (20) 204703
- 107Wang, Y.; Markwick, P. R. L.; de Oliveira, C. A. F.; McCammon, J. A. Enhanced lipid diffusion and mixing in accelerated molecular dynamics J. Chem. Theory Comput. 2011, 7 (10) 3199– 3207[ACS Full Text
], [CAS], Google Scholar
107https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFCmtL7K&md5=e170125dc7540beccc910d1d542a4247Enhanced Lipid Diffusion and Mixing in Accelerated Molecular DynamicsWang, Yi; Markwick, Phineus R. L.; de Oliveira, Cesar Augusto F.; McCammon, J. AndrewJournal of Chemical Theory and Computation (2011), 7 (10), 3199-3207CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Accelerated mol. dynamics (aMD) is an enhanced sampling technique that expedites conformational space sampling by reducing the barriers sepg. various low-energy states of a system. Here, the authors present the first application of the aMD method on lipid membranes. Altogether, ∼1.5 μs simulations were performed on three systems: a pure POPC bilayer, a pure DMPC bilayer, and a mixed POPC:DMPC bilayer. Overall, the aMD simulations are found to produce significant speedup in trans-gauche isomerization and lipid lateral diffusion vs. those in conventional MD (cMD) simulations. Further comparison of a 70-ns aMD run and a 300-ns cMD run of the mixed POPC:DMPC bilayer shows that the two simulations yield similar lipid mixing behaviors, with aMD generating a 2-3-fold speedup compared to cMD. The authors' results demonstrate that the aMD method is an efficient approach for the study of bilayer structural and dynamic properties. On the basis of simulations of the three bilayer systems, the authors also discuss the impact of aMD parameters on various lipid properties, which can be used as a guideline for future aMD simulations of membrane systems. - 108Basconi, J. E.; Shirts, M. R. Effects of temperature control algorithms on transport properties and kinetics in molecular dynamics simulations J. Chem. Theory Comput. 2013, 9 (7) 2887– 2899[ACS Full Text
], [CAS], Google Scholar
108https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXnvVKhsbg%253D&md5=d3cd49b892dc363f4177d8eca8aaad34Effects of Temperature Control Algorithms on Transport Properties and Kinetics in Molecular Dynamics SimulationsBasconi, Joseph E.; Shirts, Michael R.Journal of Chemical Theory and Computation (2013), 9 (7), 2887-2899CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Temp. control algorithms in mol. dynamics (MD) simulations are necessary to study isothermal systems. However, these thermostatting algorithms alter the velocities of the particles and thus modify the dynamics of the system with respect to the microcanonical ensemble, which could potentially lead to thermostat-dependent dynamical artifacts. In this study, we investigate how six well-established thermostat algorithms applied with different coupling strengths and to different degrees of freedom affect the dynamics of various mol. systems. We consider dynamic processes occurring on different times scales by measuring translational and rotational self-diffusion as well as the shear viscosity of water, diffusion of a small mol. solvated in water, and diffusion and the dynamic structure factor of a polymer chain in water. All of these properties are significantly dampened by thermostat algorithms which randomize particle velocities, such as the Andersen thermostat and Langevin dynamics, when strong coupling is used. For the solvated small mol. and polymer, these dampening effects are reduced somewhat if the thermostats are applied to the solvent alone, such that the solute's temp. is maintained only through thermal contact with solvent particles. Algorithms which operate by scaling the velocities, such as the Berendsen thermostat, the stochastic velocity rescaling approach of Bussi and co-workers, and the Nose-Hoover thermostat, yield transport properties that are statistically indistinguishable from those of the microcanonical ensemble, provided they are applied globally, i.e. coupled to the system's kinetic energy. When coupled to local kinetic energies, a velocity scaling thermostat can have dampening effects comparable to a velocity randomizing method, as we observe when a massive Nose-Hoover coupling scheme is used to simulate water. Correct dynamical properties, at least those studied in this paper, are obtained with the Berendsen thermostat applied globally, despite the fact that it yields the wrong kinetic energy distribution. - 109König, S.; Bayerl, T. M.; Coddens, G.; Richter, D.; Sackmann, E. Hydration dependence of chain dynamics and local diffusion in L-alpha-dipalmitoylphosphtidylcholine multilayers studied by incoherent quasi-elastic neutron scattering Biophys. J. 1995, 68 (5) 1871– 1880
Supporting Information
Supporting Information
ARTICLE SECTIONSDetails of the Lipid14 atom types, partial charges, and force field parameters. Also included are the bilayer results for additional GPU and CPU runs. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.