Transient Formation and Reactivity of a High-Valent Nickel(IV) Oxido ComplexClick to copy article linkArticle link copied!
- Sandeep K. Padamati
- Davide Angelone
- Apparao Draksharapu
- Gloria Primi
- David J. Martin
- Moniek Tromp
- Marcel Swart
- Wesley R. Browne
Abstract
A reactive high-valent dinuclear nickel(IV) oxido bridged complex is reported that can be formed at room temperature by reaction of [(L)2Ni(II)2(μ-X)3]X (X = Cl or Br) with NaOCl in methanol or acetonitrile (where L = 1,4,7-trimethyl-1,4,7-triazacyclononane). The unusual Ni(IV) oxido species is stabilized within a dinuclear tris-μ-oxido-bridged structure as [(L)2Ni(IV)2(μ-O)3]2+. Its structure and its reactivity with organic substrates are demonstrated through a combination of UV–vis absorption, resonance Raman, 1H NMR, EPR, and X-ray absorption (near-edge) spectroscopy, ESI mass spectrometry, and DFT methods. The identification of a Ni(IV)-O species opens opportunities to control the reactivity of NaOCl for selective oxidations.
Introduction
Scheme 1
Results and Discussion
Figure 1
Figure 1. Spin-density plot (S12g/TZ2P) for (antiferromagnetically coupled) 1 with spin-up spin-density shown in blue (around Nia, left) and spin-down spin-density in red (around Nib, right).
Figure 2
Figure 2. UV–vis absorption spectrum of 1 (3.5 mM) in methanol before (black) and after (red 6 s, blue 48 s, green 463 s) addition of 11 equiv of NaOCl(aq) at 293 K. Inset: Expansion of the NIR region. The data indicate that for 3 ε612 nm > 715 M–1·cm–1.
Figure 3
Scheme 2
Scheme 3
Figure 4
Figure 4. Raman spectra (λexc 532 nm) of 1 (3.5 mM) in acetonitrile (a) and with (b) 4.5 equiv of NaOBr, (c) 4.5 equiv of Na16OCl, and (d) 4.5 equiv of Na18OCl. *Solvent band. #Raman band from quartz.
Figure 5
Figure 5. Top: Calculated spectra for 3 with various degrees of 18O substitution (16O3 (blue), 16O218O (red), 16O18O2 (green), 18O3 (purple)). Bottom: Resonance Raman spectra of 3 generated from 1 (4 mM) in methanol by addition of 4 equiv of NaOCl/H2O with (blue) 100% 16O, (red) 50% 18O, (green) 34% 18O, and (purple) 26% 18O.
Conclusions
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b04158.
Animated gifs for selected vibrational modes between 517 and 893 cm−1 of 3 (ZIP)
Animated gifs for selected vibrational modes between 536 and 892 cm−1 of 3a (ZIP)
Animated gifs for vibrational modes at 653, 689, 700, 742 cm−1 of 3b (ZIP)
Details of synthesis and charcaterization of complexes 1, 2, and 4, UV–vis absorption, (resonance) Raman, NMR, EPR spectroscopy, ESI-MS, XANES, XES, and computational data (PDF)
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
The Ubbo Emmius fund of the University of Groningen, the European Research Council (StG, no. 279549, W.R.B.), NWO for a VIDI grant (723.014.010, D.J.M. and M.T.), The Netherlands Ministry of Education, Culture and Science (Gravity program 024.001.035, W.R.B.), MINECO (CTQ2014-59212-P and CTQ2015-70851-ERC, M.S.), GenCat (2014SGR1202, M.S.), FEDER (UNGI10-4E-801, M.S.), and COST action CM1305 “ECOSTBio” (W.R.B., COST-STSM-CM1305-29045) are acknowledged for financial support. A. van Dam (ERIBA) is thanked for assistance with mass spectrometry.
References
This article references 27 other publications.
- 1(a) Costas, M.; Mehn, M. P.; Jensen, M. P.; Que, L., Jr. Chem. Rev. 2004, 104, 939– 986 DOI: 10.1021/cr020628nGoogle Scholar1aDioxygen Activation at Mononuclear Nonheme Iron Active Sites: Enzymes, Models, and IntermediatesCostas, Miquel; Mehn, Mark P.; Jensen, Michael P.; Que, Lawrence, Jr.Chemical Reviews (Washington, DC, United States) (2004), 104 (2), 939-986CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The structural biol., mechanistic enzymol., phys. methods, and synthetic chem. information regarding mononuclear nonheme iron oxygenases has been complied and analyzed to demonstrate the powerful synergy among these approaches in enhancing our comprehension of how the iron centers in these enzymes function in dioxygen activation.(b) Solomon, E. I.; Brunold, T. C.; Davis, M. I.; Kemsley, J. N.; Lee, S.-K.; Lehnert, N.; Neese, F.; Skulan, A. J.; Yang, Y.-S.; Zhou, J. Chem. Rev. 2000, 100, 235– 350 DOI: 10.1021/cr9900275Google Scholar1bGeometric and Electronic Structure/Function Correlations in Non-Heme Iron EnzymesSolomon, Edward I.; Brunold, Thomas C.; Davis, Mindy I.; Kemsley, Jyllian N.; Lee, Sang-Kyu; Lehnert, Nicolai; Neese, Frank; Skulan, Andrew J.; Yang, Yi-Shan; Zhou, JingChemical Reviews (Washington, D. C.) (2000), 100 (1), 235-349CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 1011 refs. Mononuclear Non-Heme Iron. This review develops the methodologies required to study ferrous (magnetic CD spectroscopy) and ferric (ESR, magnetic CD, resonance Raman, and electronic absorption spectroscopies) active sites and then applies these to obtain geometric and electronic structural insight into the reactions of many mononuclear non-heme iron enzymes. On the basis of these studies, a general mechanistic strategy utilized by many of the non-heme ferrous enzymes has developed which is summarized. The authors then describe the new electronic structural features assocd. with binuclear iron sites with bridging ligation and consider these contributions to the reactivity of the enzymes. Geometric and electronic structure contributions to the dioxygen reactivity of the binuclear ferrous enzymes are summarized. As a general theme of this review, the authors consider what is presently known about oxygen intermediates in the proteins and related non-heme iron model complexes. The focus is on possible parallels to O2 activation by heme systems, which are thought to involve heterolytic cleavage of the O-O bond of peroxide to generate a high-valent iron-oxo intermediate (compd. I; FeIV=O, porphyrin radical), and whether such a high-valent iron-oxo intermediate would be stable in the different non-heme protein environments.(c) Punniyamurthy, T.; Velusamy, S.; Iqbal, J. Chem. Rev. 2005, 105, 2329– 2363 DOI: 10.1021/cr050523vGoogle Scholar1cRecent Advances in Transition Metal Catalyzed Oxidation of Organic Substrates with Molecular OxygenPunniyamurthy, T.; Velusamy, Subbarayan; Iqbal, JavedChemical Reviews (Washington, DC, United States) (2005), 105 (6), 2329-2363CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Included in the review are epoxidn., Wacker oxidn., allylic oxidn., etc.
- 2(a) Decker, A.; Solomon, E. I. Curr. Opin. Chem. Biol. 2005, 9, 152– 163 DOI: 10.1016/j.cbpa.2005.02.012Google Scholar2aDioxygen activation by copper, heme and non-heme iron enzymes: comparison of electronic structures and reactivitiesDecker, Andrea; Solomon, Edward I.Current Opinion in Chemical Biology (2005), 9 (2), 152-163CODEN: COCBF4; ISSN:1367-5931. (Elsevier Ltd.)A review and discussion. Enzymes contg. heme, non-heme Fe, and Cu active sites play important roles in the activation of O2 for substrate oxidn. One key reaction step is C-H bond cleavage through H-atom abstraction. On the basis of the ligand environment and the redox properties of the metal, these enzymes employ different methods of O2 activation. Heme enzymes are able to stabilize the very reactive Fe(IV)-oxo porphyrin-radical intermediate. This is generally not accessible for non-heme Fe systems, which can instead use low-spin ferric-hydroperoxo and Fe(IV)-oxo species as reactive oxidants. Cu enzymes employ still a different strategy and achieve H-atom abstraction potentially through a superoxo intermediate. This review compares and contrasts the electronic structures and reactivities of these various O intermediates.(b) Bertini, I.; Gray, H. B.; Stiefel, E. I.; Valentine, J. S. Biological.Inorganic Chemistry. Structure & Reactivity; University Science Books: Sausalito, CA, 2007.Google ScholarThere is no corresponding record for this reference.(c) Meunier, B. Chem. Rev. 1992, 92, 1411– 1456 DOI: 10.1021/cr00014a008Google ScholarThere is no corresponding record for this reference.(d) Nam, W. Acc. Chem. Res. 2015, 48, 2415– 2423 DOI: 10.1021/acs.accounts.5b00218Google Scholar2dSynthetic Mononuclear Nonheme Iron-Oxygen IntermediatesNam, WonwooAccounts of Chemical Research (2015), 48 (8), 2415-2423CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Mononuclear nonheme iron-oxygen species, such as iron-superoxo, -peroxo, -hydroperoxo, and -oxo, are key intermediates involved in dioxygen activation and oxidn. reactions catalyzed by nonheme iron enzymes. Because these iron-oxygen intermediates are short-lived due to their thermal instability and high reactivity, it is challenging to study their structural and spectroscopic properties and reactivity in the catalytic cycles of the enzymic reactions themselves. One way to approach such problems is to synthesize biomimetic iron-oxygen complexes and to tune their geometric and electronic structures for structural characterization and reactivity studies. Indeed, a no. of biol. important iron-oxygen species, such as mononuclear nonheme iron(III)-superoxo, iron(III)-peroxo, iron(III)-hydroperoxo, iron(IV)-oxo, and iron(V)-oxo complexes, were synthesized recently, and the 1st x-ray crystal structures of iron(III)-superoxo, iron(III)-peroxo, and iron(IV)-oxo complexes in nonheme iron models were successfully obtained. Thus, the authors' understanding of iron-oxygen intermediates in biol. reactions was aided greatly from the studies of the structural and spectroscopic properties and the reactivities of the synthetic biomimetic analogs. In this Account, the authors describe the authors' recent results on the synthesis and characterization of mononuclear nonheme iron-oxygen complexes bearing simple macrocyclic ligands, such as N-tetramethylated cyclam ligand (TMC) and tetraamido macrocyclic ligand (TAML). In the case of iron-superoxo complexes, an iron(III)-superoxo complex, [(TAML)FeIII(O2)]2-, is described, including its crystal structure and reactivities in electrophilic and nucleophilic oxidative reactions, and its properties are compared with those of a chromium(III)-superoxo complex, [(TMC)CrIII(O2)(Cl)]+, with respect to its reactivities in hydrogen atom transfer (HAT) and oxygen atom transfer (OAT) reactions. In the case of iron-peroxo intermediates, an x-ray crystal structure of an iron(III)-peroxo complex binding the peroxo ligand in a side-on (η2) fashion, [(TMC)FeIII(O2)]+, is described. Iron(III)-peroxo complexes binding redox-inactive metal ions are described and discussed in light of the role of redox-inactive metal ions in O-O bond activation in cytochrome c oxidase and O2-evolution in photosystem II. In the case of iron-hydroperoxo intermediates, mononuclear nonheme iron(III)-hydroperoxo complexes can be generated upon protonation of iron(III)-peroxo complexes or by hydrogen atom abstraction (HAA) of hydrocarbon C-H bonds by iron(III)-superoxo complexes. Reactivities of the iron(III)-hydroperoxo complexes in both electrophilic and nucleophilic oxidative reactions are described along with a discussion of O-O bond cleavage mechanisms. In the last section of this Account, a brief summary is presented of developments in mononuclear nonheme iron(IV)-oxo complexes since the 1st structurally characterized iron(IV)-oxo complex, [(TMC)FeIV(O)]2+, is reported. Although the field of nonheme iron-oxygen intermediates (e.g., Fe-O2, Fe-O2H, and Fe-O) was developed greatly through intense synthetic, structural, spectroscopic, reactivity, and theor. studies in the communities of bio inorg. and biomimetic chem. over the past 10 years, there is still much to be explored in trapping, characterizing, and understanding the chem. properties of the key iron-oxygen intermediates involved in dioxygen activation and oxidn. reactions by nonheme iron enzymes and their biomimetic compds.(e) Oloo, W. N.; Que, L. Acc. Chem. Res. 2015, 48, 2612– 2621 DOI: 10.1021/acs.accounts.5b00053Google ScholarThere is no corresponding record for this reference.
- 3Shearer, J. Acc. Chem. Res. 2014, 47, 2332– 2341 DOI: 10.1021/ar500060sGoogle ScholarThere is no corresponding record for this reference.
- 4(a) Cho, J.; Kang, Y.; Liu, L. V.; Sarangi, R.; Solomon, E. I.; Nam, W. Chem. Sci. 2013, 4, 1502– 1508 DOI: 10.1039/c3sc22173cGoogle Scholar4aMononuclear nickel(II)-superoxo and nickel(III)-peroxo complexes bearing a common macrocyclic TMC ligandCho, Jaeheung; Kang, Hye Yeon; Liu, Lei V.; Sarangi, Ritimukta; Solomon, Edward I.; Nam, WonwooChemical Science (2013), 4 (4), 1502-1508CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Mononuclear metal-dioxygen adducts, such as metal-superoxo and -peroxo species, were generated as key intermediates in the catalytic cycles of dioxygen activation by heme and non-heme metalloenzymes. The authors showed recently that the geometric and electronic structure of the Ni-O2 core in [Ni(n-TMC)(O2)]+ (n = 12 and 14) varies depending on the ring size of the supporting TMC ligand. Mononuclear Ni(II)-superoxo and Ni(III)-peroxo complexes bearing a common macrocyclic 13-TMC ligand, such as [NiII(13-TMC)(O2)]+ and [NiIII(13-TMC)(O2)]+, were synthesized in the reaction of [NiII(13-TMC)(MeCN)]2+ and H2O2 in the presence of tetramethylammonium hydroxide (TMAH) and triethylamine (TEA), resp. The Ni(II)-superoxo and Ni(III)-peroxo complexes bearing the common 13-TMC ligand were successfully characterized by various spectroscopic methods, x-ray crystallog. and DFT calcns. Based on the combined exptl. and theor. studies, the superoxo ligand in [NiII(13-TMC)(O2)]+ is bound in an end-on fashion to the nickel(II) center, whereas the peroxo ligand in [NiIII(13-TMC)(O2)]+ is bound in a side-on fashion to the nickel(III) center. Reactivity studies performed with the Ni(II)-superoxo and Ni(III)-peroxo complexes toward org. substrates reveal that the former possesses an electrophilic character, whereas the latter is an active oxidant in nucleophilic reaction.(b) Honda, K.; Cho, J.; Matsumoto, T.; Roh, J.; Furutachi, H.; Tosha, T.; Kubo, M.; Fujinami, S.; Ogura, T.; Kitagawa, T.; Suzuki, M. Angew. Chem., Int. Ed. 2009, 48, 3304– 3307 DOI: 10.1002/anie.200900222Google ScholarThere is no corresponding record for this reference.(c) Tano, T.; Doi, Y.; Inosako, M.; Kunishita, A.; Kubo, M.; Ishimaru, H.; Ogura, T.; Sugimoto, H.; Itoh, S. Bull. Chem. Soc. Jpn. 2010, 83, 530– 538 DOI: 10.1246/bcsj.20090346Google ScholarThere is no corresponding record for this reference.(d) Kunishita, A.; Doi, Y.; Kubo, M.; Ogura, T.; Sugimoto, H.; Itoh, S. Inorg. Chem. 2009, 48, 4997– 5004 DOI: 10.1021/ic900059mGoogle ScholarThere is no corresponding record for this reference.(e) Morimoto, Y.; Bunno, S.; Fujieda, N.; Sugimoto, H.; Itoh, S. J. Am. Chem. Soc. 2015, 137, 5867– 5870 DOI: 10.1021/jacs.5b01814Google ScholarThere is no corresponding record for this reference.
- 5(a) Corona, T.; Pfaff, F. F.; Acua-Pares, F.; Draksharapu, A.; Whiteoak, C. J.; Martin-Diaconescu, V.; Lloret-Fillol, J.; Browne, W. R.; Ray, K.; Company, A. Chem. - Eur. J. 2015, 21, 15029– 15038 DOI: 10.1002/chem.201501841Google Scholar5aReactivity of a Nickel(II) Bis(amidate) Complex with meta-Chloroperbenzoic Acid: Formation of a Potent Oxidizing SpeciesCorona, Teresa; Pfaff, Florian F.; Acuna-Pares, Ferran; Draksharapu, Apparao; Whiteoak, Christopher J.; Martin-Diaconescu, Vlad; Lloret-Fillol, Julio; Browne, Wesley R.; Ray, Kallol; Company, AnnaChemistry - A European Journal (2015), 21 (42), 15029-15038CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, we report the formation of a highly reactive nickel-oxygen species that has been trapped following reaction of a NiII precursor bearing a macrocyclic bis(amidate) ligand with meta-chloroperbenzoic acid (HmCPBA). This compd. is only detectable at temps. below 250 K and is much more reactive toward org. substrates (i.e., C-H bonds, C-C bonds, and sulfides) than previously reported well-defined nickel-oxygen species. Remarkably, this species is formed by heterolytic O-O bond cleavage of a Ni-HmCPBA precursor, which is concluded from exptl. and computational data. On the basis of spectroscopy and DFT calcns., this reactive species is proposed to be a NiIII-oxyl compd.(b) Pfaff, F. F.; Heims, F.; Kundu, S.; Mebs, S.; Ray, K. Chem. Commun. 2012, 48, 3730– 3732 DOI: 10.1039/c2cc30716bGoogle ScholarThere is no corresponding record for this reference.
- 6(a) Draksharapu, A.; Codolá, Z.; Gómez, L.; Lloret-Fillol, J.; Browne, W. R.; Costas, M. Inorg. Chem. 2015, 54, 10656– 10666 DOI: 10.1021/acs.inorgchem.5b01463Google ScholarThere is no corresponding record for this reference.(b) Pirovano, P.; Farquhar, E. R.; Swart, M.; McDonald, A. R. J. Am. Chem. Soc. 2016, 138, 14362– 14370 DOI: 10.1021/jacs.6b08406Google ScholarThere is no corresponding record for this reference.(c) Corona, T.; Draksharapu, A.; Padamati, S. K.; Gamba, I.; Martin-Diaconescu, V.; Acuña-Parés, F.; Browne, W. R.; Company, A. J. Am. Chem. Soc. 2016, 138, 12987– 12996 DOI: 10.1021/jacs.6b07544Google ScholarThere is no corresponding record for this reference.
- 7Corona, T.; Company, A. Chem. - Eur. J. 2016, 22, 13422– 13429 DOI: 10.1002/chem.201602414Google Scholar7Spectroscopically Characterized Synthetic Mononuclear Nickel-Oxygen SpeciesCorona, Teresa; Company, AnnaChemistry - A European Journal (2016), 22 (38), 13422-13429CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; iron, copper, and manganese are the predominant metals found in oxygenases that perform efficient and selective hydrocarbon oxidns. and for this reason, a large no. of the corresponding metal-oxygen species has been described. However, in recent years nickel has been found in the active site of enzymes involved in oxidn. processes, in which nickel-dioxygen species are proposed to play a key role. Owing to this biol. relevance and to the existence of different catalytic protocols that involve the use of nickel catalysts in oxidn. reactions, there is a growing interest in the detection and characterization of nickel-oxygen species relevant to these processes. In this Minireview the spectroscopically/structurally characterized synthetic superoxo, peroxo, and oxonickel species that have been reported to date are described. From these studies it becomes clear that nickel is a very promising metal in the field of oxidn. chem. with still unexplored possibilities.
- 8(a) Camasso, N. M.; Sanford, M. S. Science 2015, 347, 1– 7 DOI: 10.1126/science.aaa4526Google ScholarThere is no corresponding record for this reference.(b) Riordan, C. G. Science 2015, 347, 1203 DOI: 10.1126/science.aaa7553Google ScholarThere is no corresponding record for this reference.
- 9(a) Gray, H. B.; Hare, C. R. Inorg. Chem. 1962, 1, 363– 368 DOI: 10.1021/ic50002a034Google ScholarThere is no corresponding record for this reference.(b) O’Halloran, K. P.; Zhao, C.; Ando, N. S.; Schultz, A. J.; Koetzle, T. F.; Piccoli, P. M. B.; Hedman, B.; Hodgson, K. O.; Bobyr, E.; Kirk, M. L.; Knottenbelt, S.; Depperman, E. C.; Stein, B.; Anderson, T. M.; Cao, R.; Geletii, Y. V.; Hardcastle, K. I.; Musaev, D. G.; Neiwert, W. A.; Fang, X.; Morokuma, K.; Wu, S.; Kögerler, P.; Hill, C. L. Inorg. Chem. 2012, 51, 7025– 7031 DOI: 10.1021/ic2008914Google ScholarThere is no corresponding record for this reference.
- 10Wieghardt, K.; Schmidt, W.; Herrmann, W.; Kueppers, H. J. Inorg. Chem. 1983, 22, 2953– 2956 DOI: 10.1021/ic00162a037Google ScholarThere is no corresponding record for this reference.
- 11Bossek, U.; Nühlen, D.; Bill, E.; Glaser, T.; Krebs, C.; Weyhermüller, T.; Wieghardt, K.; Lengen, M.; Trautwein, A. X. Inorg. Chem. 1997, 36, 2834– 2843 DOI: 10.1021/ic970119hGoogle ScholarThere is no corresponding record for this reference.
- 12(a) Klamt, A.; Schüürmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 799– 805 DOI: 10.1039/P29930000799Google Scholar12aCOSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradientKlamt, A.; Schueuermann, G.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1993), (5), 799-805CODEN: JCPKBH; ISSN:0300-9580.Starting from the screening in conductors, an algorithm for the accurate calcn. of dielec. screening effects in solvents is presented, which leads to rather simple explicit expressions for the screening energy and its analytic gradient with respect to the solute coordinates. Thus geometry optimization of a solute within a realistic dielec. continuum model becomes practicable for the first time. The algorithm is suited for mol. mechanics as well as for any MO algorithm. The implementation into MOPAC and some example applications are reported.(b) Pye, C. C.; Ziegler, T. Theor. Chem. Acc. 1999, 101, 396– 408 DOI: 10.1007/s002140050457Google Scholar12bAn implementation of the conductor-like screening model of solvation within the Amsterdam density functional packagePye, Cory C.; Ziegler, TomTheoretical Chemistry Accounts (1999), 101 (6), 396-408CODEN: TCACFW; ISSN:1432-881X. (Springer-Verlag)The conductor-like screening model (COSMO) of solvation was implemented in the Amsterdam d. functional program with max. flexibility in mind. Four cavity definitions were incorporated. Several iterative schemes were tested for solving the COSMO equations. The biconjugate gradient method proves to be both robust and memory-conserving. The interaction between the surface charges and the electron d. may be calcd. by integrating over either the fitted or exact d., or by calcg. the mol. potential. A disk-smearing algorithm is applied in the former case to avoid singularities. Several SCF/COSMO coupling schemes were examd. in an attempt to reduce computational effort. A gradient-preserving algorithm for removing outlying charge was implemented. Preliminary optimized radii are given. Applications to the benzene oxide-oxepin valence tautomerization and to glycine conformation are presented.(c) Swart, M.; Rösler, E.; Bickelhaupt, F. M. Eur. J. Inorg. Chem. 2007, 2007, 3646– 3654 DOI: 10.1002/ejic.200700228Google ScholarThere is no corresponding record for this reference.
- 13van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597– 4610 DOI: 10.1063/1.466059Google Scholar13Relativistic regular two-component Hamiltoniansvan Lenthe, E.; Baerends, E. J.; Snijders, J. G.Journal of Chemical Physics (1993), 99 (6), 4597-610CODEN: JCPSA6; ISSN:0021-9606.In the present work, potential-dependent transformations were used to transform the four-component Dirac Hamiltonian into relativistic, effective, two-component, regular Hamiltonians. To zeroth order, the expansions give second-order differential equations (just like the Schroedinger equation), which already contain the most important relativistic effects, including spin-orbit coupling. One of the zeroth- order Hamiltonians is identical to the one obtained earlier by Ch. Chang, et al., (1986). By using these Hamiltonians, self-consistent all-electron and frozen-core calcns., as well as first-order perturbation calcns. were done for the uranium atom. They gave very accurate results, esp. for the one-electron energies and electron densities of the valence orbitals.
- 14(a) Swart, M. Chem. Phys. Lett. 2013, 580, 166– 171 DOI: 10.1016/j.cplett.2013.06.045Google ScholarThere is no corresponding record for this reference.(b) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098– 3100 DOI: 10.1103/PhysRevA.38.3098Google Scholar14bDensity-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.(c) Perdew, J. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33, 8822– 8824 DOI: 10.1103/PhysRevB.33.8822Google Scholar14cDensity-functional approximation for the correlation energy of the inhomogeneous electron gasPerdewPhysical review. B, Condensed matter (1986), 33 (12), 8822-8824 ISSN:0163-1829.There is no expanded citation for this reference.
- 16Tak, H.; Lee, H.; Kang, J.; Cho, J. Inorg. Chem. Front. 2016, 3, 157– 163 DOI: 10.1039/C5QI00206KGoogle Scholar16A high-spin nickel(II) borohydride complex in dehalogenationTak, Hyeonwoo; Lee, Hyunjoo; Kang, Joongoo; Cho, JaeheungInorganic Chemistry Frontiers (2016), 3 (1), 157-163CODEN: ICFNAW; ISSN:2052-1553. (Royal Society of Chemistry)A nickel(II)-borohydride complex bearing a macrocyclic tridentate N-donor ligand, [Ni(Me3-TACN)(BH4)(CH3CN)]+ (Me3-TACN = 1,4,7-trimethyl-1,4,7-triazacyclononane), was prepd., isolated, and characterized by various physicochem. methods, including UV-vis, ESI-MS, IR and X-ray analyses. The structural and spectroscopic characterization clearly shows that the borohydride ligand is bound to the high-spin nickel(II) center in an η2-manner. D. functional theory calcns. provided geometric information of 2, showing that the η2-binding of borohydride to the nickel center is more favorable than the η3-binding mode in CH3CN. The complex is paramagnetic with an effective magnetic moment of 2.9μB consistent with a d8 high-spin system. The reactivity of the high-spin nickel(II)-borohydride complex was examd. in dehalogenation with numerous halocarbons. A kinetic isotope effect value of 1.7 was obsd. in the dehalogenation of CHCl3 by the nickel(II)-borohydride complex. Kinetic studies and isotopic labeling expts. implicate that hydride ion or hydrogen atom transfer from the borohydride group is the rate detg. step. The pos. Hammett ρ value of 1.2, obtained in the reactions of [Ni(Me3-TACN)(BH4)(CH3CN)]+ and para-substituted benzoyl chloride, indicates that the dehalogenation by the nickel(II)-borohydride species occurs via a nucleophilic reaction.
- 17(a) Sun, Y.-K; Kim, M. G.; Kang, S. H.; Amine, K. J. Mater. Chem. 2003, 13, 319– 322 DOI: 10.1039/b209379kGoogle ScholarThere is no corresponding record for this reference.(b) Gu, W.; Wang, H.; Wang, K. Dalton Trans. 2014, 43, 6406– 6413 DOI: 10.1039/c4dt00308jGoogle ScholarThere is no corresponding record for this reference.
- 18Rehr, J. J.; Kas, J. J.; Vila, F. D.; Prange, M. P.; Jorissen, K. Phys. Chem. Chem. Phys. 2010, 12, 5503– 5513 DOI: 10.1039/b926434eGoogle Scholar18Parameter-free calculations of X-ray spectra with FEFF9Rehr, John J.; Kas, Joshua J.; Vila, Fernando D.; Prange, Micah P.; Jorissen, KevinPhysical Chemistry Chemical Physics (2010), 12 (21), 5503-5513CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A review. The authors briefly review implementation of the real-space Green's function (RSGF) approach for calcns. of x-ray spectra, focusing on recently developed parameter free models for dominant many-body effects. Although the RSGF approach was widely used both for near edge (XANES) and extended (EXAFS) ranges, previous implementations relied on semi-phenomenol. methods, e.g., the plasmon-pole model for the self-energy, the final-state rule for screened core hole effects, and the correlated Debye model for vibrational damping. Here the authors describe how these approxns. can be replaced by efficient ab initio models including a many-pole model of the self-energy, inelastic losses and multiple-electron excitations; a linear response approach for the core hole; and a Lanczos approach for Debye-Waller effects. The authors also discuss the implementation of these models and software improvements within the FEFF9 code, together with a no. of examples.
- 19
In the absence of Ni(II) or 1, NaOCl is stable in CH3CN.
There is no corresponding record for this reference. - 20Inosako, M.; Kunishita, A.; Kubo, M.; Ogura, T.; Sugimoto, T.; Itoh, S. Dalton Trans. 2009, 43, 9410– 9147 DOI: 10.1039/b910237j
This complex is reminiscent of the (μ-η2:η2-disulfido)dinickel(II) complexes with 6-methyl-TPA ligands reported by Itoh et al.; see:
Google ScholarThere is no corresponding record for this reference. - 21
The formation of species such as (O═Ni(IV)–O–Ni(IV)═O) is highly unlikely, as it is calculated to lie 45 to 58 kcal mol–1 (depending on functional used, e.g., BP86-D3 and S12g) higher in energy than even the peroxy species 3a. Furthermore, DFT calculations indicate a high tendency for a species to converge to the structure of the 3a species. In addition ligand hydroxylation can be discounted on the basis of the recovery of the Ni(II) complexes, e.g., 1 and 4, once 3 has reacted with solvent or substrate.
There is no corresponding record for this reference. - 22Wieghardt, K.; Bossek, U.; Nuber, B.; Weiss, J.; Bonvoisin, J.; Corbella, M.; Vitols, S. E.; Girerd, J. J. J. Am. Chem. Soc. 1988, 110, 7398– 7411 DOI: 10.1021/ja00230a021Google ScholarThere is no corresponding record for this reference.
- 23(a) Hage, R.; Krijnen, B.; Warnaar, J. B.; Hartl, F.; Stufkens, D. J.; Snoeck, T. L. Inorg. Chem. 1995, 34, 4973– 4978 DOI: 10.1021/ic00124a010Google ScholarThere is no corresponding record for this reference.(b) Angelone, D.; Abdolahzadeh, S.; de Boer, J. W.; Browne, W. R. Eur. J. Inorg. Chem. 2015, 21, 3532– 3542 DOI: 10.1002/ejic.201500195Google ScholarThere is no corresponding record for this reference.
- 24Mandimutsira, B. S.; Yamarik, J. L.; Brunold, T. C.; Gu, W.; Cramer, S. P.; Riordan, C. G. J. Am. Chem. Soc. 2001, 123, 9194– 9195 DOI: 10.1021/ja016209+Google ScholarThere is no corresponding record for this reference.
PhTttBu = phenyltris((tert-butylthio)methyl)borate.
- 25Itoh, S.; Bandoh, H.; Nakagawa, M.; Nagatomo, S.; Kitagawa, T.; Karlin, K. D.; Fukuzumi, S. J. Am. Chem. Soc. 2001, 123, 11168– 11178 DOI: 10.1021/ja0104094Google ScholarThere is no corresponding record for this reference.
- 26Kieber-Emmons, M. T.; Schenker, R.; Yap, G. P. A.; Brunold, T. C.; Riordan, C. G. Angew. Chem. 2004, 116, 6884– 6886 DOI: 10.1002/ange.200460747Google ScholarThere is no corresponding record for this reference.
tmc = 1,4,8,11-tetramethyl-1,4,8,11-tetraazadodecane.
- 27Rettenmeier, C. A.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2015, 54, 4880– 4884 DOI: 10.1002/anie.201500141Google ScholarThere is no corresponding record for this reference.
iso-pmbox = bis(oxazolinylmethylidene)pyrrolidine.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 51 publications.
- Avery LeComte, Rachel Sailer, Samyadeb Mahato, Warren VandeVen, Wen Zhou, Alisa R. Paterson, Morgane Desmau, Amani M. Ebrahim, Fabrice Thomas, Linus Chiang. Synthesis and Characterization of Co Complexes Coordinated by a Tetraanionic Bis(amidateanilido) Ligand. Inorganic Chemistry 2025, 64
(12)
, 5986-5995. https://doi.org/10.1021/acs.inorgchem.4c05005
- Deepika G. Karmalkar, Hyeongtaek Lim, Mahesh Sundararajan, Yong-Min Lee, Mi Sook Seo, Dae Young Bae, Xiaoyan Lu, Britt Hedman, Keith O. Hodgson, Won-Suk Kim, Eunsung Lee, Edward I. Solomon, Shunichi Fukuzumi, Wonwoo Nam. Synthesis, Structure, and Redox Reactivity of Ni Complexes Bearing a Redox and Acid–Base Non-innocent Ligand with NiII, NiIII, and NiIV Formal Oxidation States. Journal of the American Chemical Society 2025, 147
(5)
, 3981-3993. https://doi.org/10.1021/jacs.4c11751
- Bach X. Nguyen, Warren VandeVen, Gregory A. MacNeil, Wen Zhou, Alisa R. Paterson, Charles J. Walsby, Linus Chiang. High-Valent Ni and Cu Complexes of a Tetraanionic Bis(amidateanilido) Ligand. Inorganic Chemistry 2023, 62
(37)
, 15180-15194. https://doi.org/10.1021/acs.inorgchem.3c02358
- Jibo Zhang, Jay R. Winkler, Harry B. Gray, Bryan M. Hunter. Mechanism of Nickel–Iron Water Oxidation Electrocatalysts. Energy & Fuels 2021, 35
(23)
, 19164-19169. https://doi.org/10.1021/acs.energyfuels.1c02674
- Ivan S. Golovanov, Roman S. Malykhin, Vladislav K. Lesnikov, Yulia V. Nelyubina, Valentin V. Novikov, Kirill V. Frolov, Andrey I. Stadnichenko, Evgeny V. Tretyakov, Sema L. Ioffe, Alexey Yu. Sukhorukov. Revealing the Structure of Transition Metal Complexes of Formaldoxime. Inorganic Chemistry 2021, 60
(8)
, 5523-5537. https://doi.org/10.1021/acs.inorgchem.0c03362
- Norman Zhao, Alexander S. Filatov, Jiaze Xie, Ethan A. Hill, Andrey Yu. Rogachev, John S. Anderson. Generation and Reactivity of a NiIII2(μ-1,2-peroxo) Complex. Journal of the American Chemical Society 2020, 142
(52)
, 21634-21639. https://doi.org/10.1021/jacs.0c10958
- Franck Le Vaillant, Edward J. Reijerse, Markus Leutzsch, Josep Cornella. Dialkyl Ether Formation at High-Valent Nickel. Journal of the American Chemical Society 2020, 142
(46)
, 19540-19550. https://doi.org/10.1021/jacs.0c07381
- Jorn D. Steen, Stepan Stepanovic, Mahsa Parvizian, Johannes W. de Boer, Ronald Hage, Juan Chen, Marcel Swart, Maja Gruden, Wesley R. Browne. Lewis versus Brønsted Acid Activation of a Mn(IV) Catalyst for Alkene Oxidation. Inorganic Chemistry 2019, 58
(21)
, 14924-14930. https://doi.org/10.1021/acs.inorgchem.9b02737
- Juan Chen, Apparao Draksharapu, Davide Angelone, Duenpen Unjaroen, Sandeep K. Padamati, Ronald Hage, Marcel Swart, Carole Duboc, Wesley R. Browne. H2O2 Oxidation by FeIII–OOH Intermediates and Its Effect on Catalytic Efficiency. ACS Catalysis 2018, 8
(10)
, 9665-9674. https://doi.org/10.1021/acscatal.8b02326
- Peng-Cheng Duan, Dennis-Helmut Manz, Sebastian Dechert, Serhiy Demeshko, Franc Meyer. Reductive O2 Binding at a Dihydride Complex Leading to Redox Interconvertible μ-1,2-Peroxo and μ-1,2-Superoxo Dinickel(II) Intermediates. Journal of the American Chemical Society 2018, 140
(14)
, 4929-4939. https://doi.org/10.1021/jacs.8b01468
- Maria
Letizia Merlini, George J. P. Britovsek, Marcel Swart, Paola Belanzoni. Understanding the Catalase-Like Activity of a Bioinspired Manganese(II) Complex with a Pentadentate NSNSN Ligand Framework. A Computational Insight into the Mechanism. ACS Catalysis 2018, 8
(4)
, 2944-2958. https://doi.org/10.1021/acscatal.7b03559
- Nicole M. Camasso, Allan J. Canty, Alireza Ariafard, and Melanie S. Sanford . Experimental and Computational Studies of High-Valent Nickel and Palladium Complexes. Organometallics 2017, 36
(22)
, 4382-4393. https://doi.org/10.1021/acs.organomet.7b00613
- Charvi Singhvi, Gunjan Sharma, Rishi Verma, Vinod K. Paidi, Pieter Glatzel, Paul Paciok, Vashishtha B. Patel, Ojus Mohan, Vivek Polshettiwar. Tuning the electronic structure and SMSI by integrating trimetallic sites with defective ceria for the CO
2
reduction reaction. Proceedings of the National Academy of Sciences 2025, 122
(3)
https://doi.org/10.1073/pnas.2411406122
- Thomas M. Khazanov, Anusree Mukherjee. Harnessing Oxidizing Potential of Nickel for Sustainable Hydrocarbon Functionalization. Molecules 2024, 29
(21)
, 5188. https://doi.org/10.3390/molecules29215188
- Ayushi Awasthi, Sharath Chandra Mallojjala, Rakesh Kumar, Raju Eerlapally, Jennifer S. Hirschi, Apparao Draksharapu. Altering the Localization of an Unpaired Spin in a
Formal
Ni(V) Species. Chemistry – A European Journal 2024, 30
(4)
https://doi.org/10.1002/chem.202302824
- Eduardo Jaimes–Romano, Hugo Valdés, Simon Hernández–Ortega, Rosa Mollfulleda, Marcel Swart, David Morales–Morales. C–S couplings catalyzed by Ni(II) complexes of the type [(NHC)Ni(Cp)(Br)]. Journal of Catalysis 2023, 426 , 247-256. https://doi.org/10.1016/j.jcat.2023.07.001
- Philipp Heim, Giuseppe Spedalotto, Marta Lovisari, Robert Gericke, John O'Brien, Erik R. Farquhar, Aidan R. McDonald. Synthesis and Characterization of a Masked Terminal Nickel‐Oxide Complex. Chemistry – A European Journal 2023, 29
(21)
https://doi.org/10.1002/chem.202203840
- Priya Sahni, Rahat Gupta, Simran Sharma, Amlan K. Pal. The once-elusive Ni(IV) species is now a potent candidate for challenging organic transformations. Coordination Chemistry Reviews 2023, 474 , 214849. https://doi.org/10.1016/j.ccr.2022.214849
- Bing-Feng Qian, Yang Gao, Qian-Ya Xu, Ai-Quan Jia, Hua-Tian Shi, Qian-Feng Zhang. Syntheses, characterizations and structures of ruthenium carbene and allenylidene complexes supported by 1,4,7-trimethyl-1,4,7-triazacyclononane (Me3tacn) ligands. Journal of Organometallic Chemistry 2022, 976 , 122429. https://doi.org/10.1016/j.jorganchem.2022.122429
- Jayanta Bag, Kuntal Pal. The access of {NiIV(OH)2} intermediate in Ni(II) mediated oxygen atom transfer to coordinated Phosphine: Combined experimental and computational studies. Polyhedron 2022, 224 , 116030. https://doi.org/10.1016/j.poly.2022.116030
- Denis V. Chachkov, Oleg V. Mikhailov. Nickel macrocyclic complexes with porphyrazine and some [benzo]substituted, oxo and fluoro ligands: DFT analysis. Journal of Porphyrins and Phthalocyanines 2022, 26
(03)
, 222-231. https://doi.org/10.1142/S1088424622500067
- Evan P. Jahrman, Jamie L. Weaver, Niranjan Govind, Marko Perestjuk, Gerald T. Seidler. Iron redox analysis of silicate-based minerals and glasses using synchrotron X-ray absorption and laboratory X-ray emission spectroscopy. Journal of Non-Crystalline Solids 2022, 577 , 121326. https://doi.org/10.1016/j.jnoncrysol.2021.121326
- Mian Guo, Yong-Min Lee, Shunichi Fukuzumi, Wonwoo Nam. Biomimetic metal-oxidant adducts as active oxidants in oxidation reactions. Coordination Chemistry Reviews 2021, 435 , 213807. https://doi.org/10.1016/j.ccr.2021.213807
- Olga V. Safonova, Maarten Nachtegaal.
Operando
X
‐Ray Spectroscopies on Catalysts in Action. 2021, 339-361. https://doi.org/10.1002/9783527813599.ch19
- Soohyung Kim, Ha Young Jeong, Seonghan Kim, Hongsik Kim, Sojeong Lee, Jaeheung Cho, Cheal Kim, Dongwhan Lee. Proton Switch in the Secondary Coordination Sphere to Control Catalytic Events at the Metal Center: Biomimetic Oxo Transfer Chemistry of Nickel Amidate Complex. Chemistry – A European Journal 2021, 27
(14)
, 4700-4708. https://doi.org/10.1002/chem.202005183
- Bing-Feng Qian, Jun-Ling Wang, Ai-Quan Jia, Hua-Tian Shi, Qian-Feng Zhang. Syntheses, reactivity, structures and photocatalytic properties of mononuclear ruthenium(II) complexes supported by 1,4,7-trimethyl-1,4,7-triazacyclononane (Me3tacn) ligands. Inorganica Chimica Acta 2021, 516 , 120128. https://doi.org/10.1016/j.ica.2020.120128
- Neil Heberer, Chi-Herng Hu, Liviu M. Mirica. High-Valent Ni Coordination Compounds. 2021, 348-374. https://doi.org/10.1016/B978-0-08-102688-5.00104-5
- Joeri Hessels, Eduard Masferrer‐Rius, Fengshou Yu, Remko J. Detz, Robertus J. M. Klein Gebbink, Joost N. H. Reek. Nickel is a Different Pickle: Trends in Water Oxidation Catalysis for Molecular Nickel Complexes. ChemSusChem 2020, 13
(24)
, 6629-6634. https://doi.org/10.1002/cssc.202002164
- Patric Zimmermann, Sergey Peredkov, Paula Macarena Abdala, Serena DeBeer, Moniek Tromp, Christoph Müller, Jeroen A. van Bokhoven. Modern X-ray spectroscopy: XAS and XES in the laboratory. Coordination Chemistry Reviews 2020, 423 , 213466. https://doi.org/10.1016/j.ccr.2020.213466
- Lorenzo D'Amore, Leonardo Belpassi, Johannes E. M. N. Klein, Marcel Swart. Spin-resolved charge displacement analysis as an intuitive tool for the evaluation of cPCET and HAT scenarios. Chemical Communications 2020, 56
(81)
, 12146-12149. https://doi.org/10.1039/D0CC04995F
- Cai-Xia Zhang, Duo-Wen Fang, Jun-Ling Wang, Ai-Quan Jia, Qian-Feng Zhang. Syntheses, characterizations, and reactivities of new 1,4,7-trimethyl-l,4,7-triazacyclononane (Me3tacn) molybdenum and tungsten complexes. Inorganica Chimica Acta 2020, 507 , 119599. https://doi.org/10.1016/j.ica.2020.119599
- Filip Vlahovic, Maja Gruden, Stepan Stepanovic, Marcel Swart. Density functional approximations for consistent spin and oxidation states of oxoiron complexes. International Journal of Quantum Chemistry 2020, 120
(5)
https://doi.org/10.1002/qua.26121
- Noel Nebra. High-Valent NiIII and NiIV Species Relevant to C–C and C–Heteroatom Cross-Coupling Reactions: State of the Art. Molecules 2020, 25
(5)
, 1141. https://doi.org/10.3390/molecules25051141
- Liviu M. Mirica, Sofia M. Smith, Leonel Griego. Organometallic Chemistry of High‐Valent
Ni
(
III
) and
Ni
(
IV
) Complexes. 2020, 223-248. https://doi.org/10.1002/9783527813827.ch10
- Marcel Swart. Dealing with Spin States in Computational Organometallic Catalysis. 2020, 191-226. https://doi.org/10.1007/3418_2020_49
- Muniyandi Sankaralingam, Mani Balamurugan, Mallayan Palaniandavar. Alkane and alkene oxidation reactions catalyzed by nickel(II) complexes: Effect of ligand factors. Coordination Chemistry Reviews 2020, 403 , 213085. https://doi.org/10.1016/j.ccr.2019.213085
- Sofia M. Smith, Oriol Planas, Laura Gómez, Nigam P. Rath, Xavi Ribas, Liviu M. Mirica. Aerobic C–C and C–O bond formation reactions mediated by high-valent nickel species. Chemical Science 2019, 10
(44)
, 10366-10372. https://doi.org/10.1039/C9SC03758F
- Giuseppe Spedalotto, Robert Gericke, Marta Lovisari, Erik R. Farquhar, Brendan Twamley, Aidan R. McDonald. Preparation and Characterisation of a Bis‐μ‐Hydroxo‐Ni
III
2
Complex. Chemistry – A European Journal 2019, 25
(51)
, 11983-11990. https://doi.org/10.1002/chem.201902812
- Xiaojun Yang, Hang Bai, Marie Moriko Qian, Changjiang Wu, Shuo Cao, Zhiguo Yan, Qifeng Tian, Zhiping Du. Enhanced activity of Pd–Cu(Ni) bimetallic catalysts for oxidative carbonylation of phenol. Materials Chemistry and Physics 2019, 234 , 48-54. https://doi.org/10.1016/j.matchemphys.2019.05.049
- José G. Moya‐Cancino, Ari‐Pekka Honkanen, Ad M. J. van der Eerden, Herrick Schaink, Lieven Folkertsma, Mahnaz Ghiasi, Alessandro Longo, Florian Meirer, Frank M. F. de Groot, Simo Huotari, Bert M. Weckhuysen. Elucidating the K‐Edge X‐Ray Absorption Near‐Edge Structure of Cobalt Carbide. ChemCatChem 2019, 11
(13)
, 3042-3045. https://doi.org/10.1002/cctc.201900434
- Miquel Costas. Alkane Oxidation with Biologically Inspired Nonheme Iron Catalysts Based in the Triazacyclononane Ligand Scaffold. 2019, 251-268. https://doi.org/10.1002/9781119379256.ch13
- Sarvesh S. Harmalkar, Dattaprasad D. Narulkar, Raymond J. Butcher, Mahesh S. Deshmukh, Anant Kumar Srivastava, Mariappan Mariappan, Prem Lama, Sunder N. Dhuri. Dual-site aqua mononuclear nickel(II) complexes of non-heme tetradentate ligands: Synthesis, characterization and reactivity. Inorganica Chimica Acta 2019, 486 , 425-434. https://doi.org/10.1016/j.ica.2018.10.069
- Evan P. Jahrman, William M. Holden, Alexander S. Ditter, Devon R. Mortensen, Gerald T. Seidler, Timothy T. Fister, Stosh A. Kozimor, Louis F. J. Piper, Jatinkumar Rana, Neil C. Hyatt, Martin C. Stennett. An improved laboratory-based x-ray absorption fine structure and x-ray emission spectrometer for analytical applications in materials chemistry research. Review of Scientific Instruments 2019, 90
(2)
https://doi.org/10.1063/1.5049383
- Maja Gruden, Wesley R. Browne, Marcel Swart, Carole Duboc. Computational Versus Experimental Spectroscopy for Transition Metals. 2019, 161-183. https://doi.org/10.1007/978-3-030-11714-6_6
- Evan P. Jahrman, Lisa A. Pellerin, Alexander S. Ditter, Liam R. Bradshaw, Timothy T. Fister, Bryant J. Polzin, Steven E. Trask, Alison R. Dunlop, Gerald T. Seidler. Laboratory-Based X-ray Absorption Spectroscopy on a Working Pouch Cell Battery at Industrially-Relevant Charging Rates. Journal of The Electrochemical Society 2019, 166
(12)
, A2549-A2555. https://doi.org/10.1149/2.0721912jes
- Bhawana Pandey, Kallol Ray, Gopalan Rajaraman. Structure, Bonding, Reactivity and Spectral Features of Putative Ni
III
=O Species: A Theoretical Perspective. Zeitschrift für anorganische und allgemeine Chemie 2018, 644
(14)
, 790-800. https://doi.org/10.1002/zaac.201800122
- Hitoshi Abe, Giuliana Aquilanti, Roberto Boada, Bruce Bunker, Pieter Glatzel, Maarten Nachtegaal, Sakura Pascarelli. Improving the quality of XAFS data. Journal of Synchrotron Radiation 2018, 25
(4)
, 972-980. https://doi.org/10.1107/S1600577518006021
- Muniyandi Sankaralingam, Yong-Min Lee, Wonwoo Nam, Shunichi Fukuzumi. Amphoteric reactivity of metal–oxygen complexes in oxidation reactions. Coordination Chemistry Reviews 2018, 365 , 41-59. https://doi.org/10.1016/j.ccr.2018.03.003
- Masahiro Kouno, Nobuto Yoshinari, Naoto Kuwamura, Kohei Yamagami, Akira Sekiyama, Mitsutaka Okumura, Takumi Konno. Valence Interconversion of Octahedral Nickel(II/III/IV) Centers. Angewandte Chemie 2017, 129
(44)
, 13950-13954. https://doi.org/10.1002/ange.201708169
- Masahiro Kouno, Nobuto Yoshinari, Naoto Kuwamura, Kohei Yamagami, Akira Sekiyama, Mitsutaka Okumura, Takumi Konno. Valence Interconversion of Octahedral Nickel(II/III/IV) Centers. Angewandte Chemie International Edition 2017, 56
(44)
, 13762-13766. https://doi.org/10.1002/anie.201708169
- Susanta Hazra, Nuno M. R. Martins, Maxim L. Kuznetsov, M. Fátima C. Guedes da Silva, Armando J. L. Pombeiro. Flexibility and lability of a phenyl ligand in hetero-organometallic 3d metal–Sn(
iv
) compounds and their catalytic activity in Baeyer–Villiger oxidation of cyclohexanone. Dalton Transactions 2017, 46
(39)
, 13364-13375. https://doi.org/10.1039/C7DT02534C
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Scheme 1
Scheme 1. Formation of 3 from 1 and NaOClFigure 1
Figure 1. Spin-density plot (S12g/TZ2P) for (antiferromagnetically coupled) 1 with spin-up spin-density shown in blue (around Nia, left) and spin-down spin-density in red (around Nib, right).
Figure 2
Figure 2. UV–vis absorption spectrum of 1 (3.5 mM) in methanol before (black) and after (red 6 s, blue 48 s, green 463 s) addition of 11 equiv of NaOCl(aq) at 293 K. Inset: Expansion of the NIR region. The data indicate that for 3 ε612 nm > 715 M–1·cm–1.
Figure 3
Scheme 2
Scheme 2. Formation of 3 from 1 and NaOCl and Subsequent Decay to 4 in AcetonitrileScheme 3
Scheme 3. Structures (3, 3a, 3b) Consistent with ESI Mass Spectral Data and Calculated Driving Forces for Their Formation from 1Figure 4
Figure 4. Raman spectra (λexc 532 nm) of 1 (3.5 mM) in acetonitrile (a) and with (b) 4.5 equiv of NaOBr, (c) 4.5 equiv of Na16OCl, and (d) 4.5 equiv of Na18OCl. *Solvent band. #Raman band from quartz.
Figure 5
Figure 5. Top: Calculated spectra for 3 with various degrees of 18O substitution (16O3 (blue), 16O218O (red), 16O18O2 (green), 18O3 (purple)). Bottom: Resonance Raman spectra of 3 generated from 1 (4 mM) in methanol by addition of 4 equiv of NaOCl/H2O with (blue) 100% 16O, (red) 50% 18O, (green) 34% 18O, and (purple) 26% 18O.
References
This article references 27 other publications.
- 1(a) Costas, M.; Mehn, M. P.; Jensen, M. P.; Que, L., Jr. Chem. Rev. 2004, 104, 939– 986 DOI: 10.1021/cr020628n1aDioxygen Activation at Mononuclear Nonheme Iron Active Sites: Enzymes, Models, and IntermediatesCostas, Miquel; Mehn, Mark P.; Jensen, Michael P.; Que, Lawrence, Jr.Chemical Reviews (Washington, DC, United States) (2004), 104 (2), 939-986CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The structural biol., mechanistic enzymol., phys. methods, and synthetic chem. information regarding mononuclear nonheme iron oxygenases has been complied and analyzed to demonstrate the powerful synergy among these approaches in enhancing our comprehension of how the iron centers in these enzymes function in dioxygen activation.(b) Solomon, E. I.; Brunold, T. C.; Davis, M. I.; Kemsley, J. N.; Lee, S.-K.; Lehnert, N.; Neese, F.; Skulan, A. J.; Yang, Y.-S.; Zhou, J. Chem. Rev. 2000, 100, 235– 350 DOI: 10.1021/cr99002751bGeometric and Electronic Structure/Function Correlations in Non-Heme Iron EnzymesSolomon, Edward I.; Brunold, Thomas C.; Davis, Mindy I.; Kemsley, Jyllian N.; Lee, Sang-Kyu; Lehnert, Nicolai; Neese, Frank; Skulan, Andrew J.; Yang, Yi-Shan; Zhou, JingChemical Reviews (Washington, D. C.) (2000), 100 (1), 235-349CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 1011 refs. Mononuclear Non-Heme Iron. This review develops the methodologies required to study ferrous (magnetic CD spectroscopy) and ferric (ESR, magnetic CD, resonance Raman, and electronic absorption spectroscopies) active sites and then applies these to obtain geometric and electronic structural insight into the reactions of many mononuclear non-heme iron enzymes. On the basis of these studies, a general mechanistic strategy utilized by many of the non-heme ferrous enzymes has developed which is summarized. The authors then describe the new electronic structural features assocd. with binuclear iron sites with bridging ligation and consider these contributions to the reactivity of the enzymes. Geometric and electronic structure contributions to the dioxygen reactivity of the binuclear ferrous enzymes are summarized. As a general theme of this review, the authors consider what is presently known about oxygen intermediates in the proteins and related non-heme iron model complexes. The focus is on possible parallels to O2 activation by heme systems, which are thought to involve heterolytic cleavage of the O-O bond of peroxide to generate a high-valent iron-oxo intermediate (compd. I; FeIV=O, porphyrin radical), and whether such a high-valent iron-oxo intermediate would be stable in the different non-heme protein environments.(c) Punniyamurthy, T.; Velusamy, S.; Iqbal, J. Chem. Rev. 2005, 105, 2329– 2363 DOI: 10.1021/cr050523v1cRecent Advances in Transition Metal Catalyzed Oxidation of Organic Substrates with Molecular OxygenPunniyamurthy, T.; Velusamy, Subbarayan; Iqbal, JavedChemical Reviews (Washington, DC, United States) (2005), 105 (6), 2329-2363CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Included in the review are epoxidn., Wacker oxidn., allylic oxidn., etc.
- 2(a) Decker, A.; Solomon, E. I. Curr. Opin. Chem. Biol. 2005, 9, 152– 163 DOI: 10.1016/j.cbpa.2005.02.0122aDioxygen activation by copper, heme and non-heme iron enzymes: comparison of electronic structures and reactivitiesDecker, Andrea; Solomon, Edward I.Current Opinion in Chemical Biology (2005), 9 (2), 152-163CODEN: COCBF4; ISSN:1367-5931. (Elsevier Ltd.)A review and discussion. Enzymes contg. heme, non-heme Fe, and Cu active sites play important roles in the activation of O2 for substrate oxidn. One key reaction step is C-H bond cleavage through H-atom abstraction. On the basis of the ligand environment and the redox properties of the metal, these enzymes employ different methods of O2 activation. Heme enzymes are able to stabilize the very reactive Fe(IV)-oxo porphyrin-radical intermediate. This is generally not accessible for non-heme Fe systems, which can instead use low-spin ferric-hydroperoxo and Fe(IV)-oxo species as reactive oxidants. Cu enzymes employ still a different strategy and achieve H-atom abstraction potentially through a superoxo intermediate. This review compares and contrasts the electronic structures and reactivities of these various O intermediates.(b) Bertini, I.; Gray, H. B.; Stiefel, E. I.; Valentine, J. S. Biological.Inorganic Chemistry. Structure & Reactivity; University Science Books: Sausalito, CA, 2007.There is no corresponding record for this reference.(c) Meunier, B. Chem. Rev. 1992, 92, 1411– 1456 DOI: 10.1021/cr00014a008There is no corresponding record for this reference.(d) Nam, W. Acc. Chem. Res. 2015, 48, 2415– 2423 DOI: 10.1021/acs.accounts.5b002182dSynthetic Mononuclear Nonheme Iron-Oxygen IntermediatesNam, WonwooAccounts of Chemical Research (2015), 48 (8), 2415-2423CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Mononuclear nonheme iron-oxygen species, such as iron-superoxo, -peroxo, -hydroperoxo, and -oxo, are key intermediates involved in dioxygen activation and oxidn. reactions catalyzed by nonheme iron enzymes. Because these iron-oxygen intermediates are short-lived due to their thermal instability and high reactivity, it is challenging to study their structural and spectroscopic properties and reactivity in the catalytic cycles of the enzymic reactions themselves. One way to approach such problems is to synthesize biomimetic iron-oxygen complexes and to tune their geometric and electronic structures for structural characterization and reactivity studies. Indeed, a no. of biol. important iron-oxygen species, such as mononuclear nonheme iron(III)-superoxo, iron(III)-peroxo, iron(III)-hydroperoxo, iron(IV)-oxo, and iron(V)-oxo complexes, were synthesized recently, and the 1st x-ray crystal structures of iron(III)-superoxo, iron(III)-peroxo, and iron(IV)-oxo complexes in nonheme iron models were successfully obtained. Thus, the authors' understanding of iron-oxygen intermediates in biol. reactions was aided greatly from the studies of the structural and spectroscopic properties and the reactivities of the synthetic biomimetic analogs. In this Account, the authors describe the authors' recent results on the synthesis and characterization of mononuclear nonheme iron-oxygen complexes bearing simple macrocyclic ligands, such as N-tetramethylated cyclam ligand (TMC) and tetraamido macrocyclic ligand (TAML). In the case of iron-superoxo complexes, an iron(III)-superoxo complex, [(TAML)FeIII(O2)]2-, is described, including its crystal structure and reactivities in electrophilic and nucleophilic oxidative reactions, and its properties are compared with those of a chromium(III)-superoxo complex, [(TMC)CrIII(O2)(Cl)]+, with respect to its reactivities in hydrogen atom transfer (HAT) and oxygen atom transfer (OAT) reactions. In the case of iron-peroxo intermediates, an x-ray crystal structure of an iron(III)-peroxo complex binding the peroxo ligand in a side-on (η2) fashion, [(TMC)FeIII(O2)]+, is described. Iron(III)-peroxo complexes binding redox-inactive metal ions are described and discussed in light of the role of redox-inactive metal ions in O-O bond activation in cytochrome c oxidase and O2-evolution in photosystem II. In the case of iron-hydroperoxo intermediates, mononuclear nonheme iron(III)-hydroperoxo complexes can be generated upon protonation of iron(III)-peroxo complexes or by hydrogen atom abstraction (HAA) of hydrocarbon C-H bonds by iron(III)-superoxo complexes. Reactivities of the iron(III)-hydroperoxo complexes in both electrophilic and nucleophilic oxidative reactions are described along with a discussion of O-O bond cleavage mechanisms. In the last section of this Account, a brief summary is presented of developments in mononuclear nonheme iron(IV)-oxo complexes since the 1st structurally characterized iron(IV)-oxo complex, [(TMC)FeIV(O)]2+, is reported. Although the field of nonheme iron-oxygen intermediates (e.g., Fe-O2, Fe-O2H, and Fe-O) was developed greatly through intense synthetic, structural, spectroscopic, reactivity, and theor. studies in the communities of bio inorg. and biomimetic chem. over the past 10 years, there is still much to be explored in trapping, characterizing, and understanding the chem. properties of the key iron-oxygen intermediates involved in dioxygen activation and oxidn. reactions by nonheme iron enzymes and their biomimetic compds.(e) Oloo, W. N.; Que, L. Acc. Chem. Res. 2015, 48, 2612– 2621 DOI: 10.1021/acs.accounts.5b00053There is no corresponding record for this reference.
- 3Shearer, J. Acc. Chem. Res. 2014, 47, 2332– 2341 DOI: 10.1021/ar500060sThere is no corresponding record for this reference.
- 4(a) Cho, J.; Kang, Y.; Liu, L. V.; Sarangi, R.; Solomon, E. I.; Nam, W. Chem. Sci. 2013, 4, 1502– 1508 DOI: 10.1039/c3sc22173c4aMononuclear nickel(II)-superoxo and nickel(III)-peroxo complexes bearing a common macrocyclic TMC ligandCho, Jaeheung; Kang, Hye Yeon; Liu, Lei V.; Sarangi, Ritimukta; Solomon, Edward I.; Nam, WonwooChemical Science (2013), 4 (4), 1502-1508CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)Mononuclear metal-dioxygen adducts, such as metal-superoxo and -peroxo species, were generated as key intermediates in the catalytic cycles of dioxygen activation by heme and non-heme metalloenzymes. The authors showed recently that the geometric and electronic structure of the Ni-O2 core in [Ni(n-TMC)(O2)]+ (n = 12 and 14) varies depending on the ring size of the supporting TMC ligand. Mononuclear Ni(II)-superoxo and Ni(III)-peroxo complexes bearing a common macrocyclic 13-TMC ligand, such as [NiII(13-TMC)(O2)]+ and [NiIII(13-TMC)(O2)]+, were synthesized in the reaction of [NiII(13-TMC)(MeCN)]2+ and H2O2 in the presence of tetramethylammonium hydroxide (TMAH) and triethylamine (TEA), resp. The Ni(II)-superoxo and Ni(III)-peroxo complexes bearing the common 13-TMC ligand were successfully characterized by various spectroscopic methods, x-ray crystallog. and DFT calcns. Based on the combined exptl. and theor. studies, the superoxo ligand in [NiII(13-TMC)(O2)]+ is bound in an end-on fashion to the nickel(II) center, whereas the peroxo ligand in [NiIII(13-TMC)(O2)]+ is bound in a side-on fashion to the nickel(III) center. Reactivity studies performed with the Ni(II)-superoxo and Ni(III)-peroxo complexes toward org. substrates reveal that the former possesses an electrophilic character, whereas the latter is an active oxidant in nucleophilic reaction.(b) Honda, K.; Cho, J.; Matsumoto, T.; Roh, J.; Furutachi, H.; Tosha, T.; Kubo, M.; Fujinami, S.; Ogura, T.; Kitagawa, T.; Suzuki, M. Angew. Chem., Int. Ed. 2009, 48, 3304– 3307 DOI: 10.1002/anie.200900222There is no corresponding record for this reference.(c) Tano, T.; Doi, Y.; Inosako, M.; Kunishita, A.; Kubo, M.; Ishimaru, H.; Ogura, T.; Sugimoto, H.; Itoh, S. Bull. Chem. Soc. Jpn. 2010, 83, 530– 538 DOI: 10.1246/bcsj.20090346There is no corresponding record for this reference.(d) Kunishita, A.; Doi, Y.; Kubo, M.; Ogura, T.; Sugimoto, H.; Itoh, S. Inorg. Chem. 2009, 48, 4997– 5004 DOI: 10.1021/ic900059mThere is no corresponding record for this reference.(e) Morimoto, Y.; Bunno, S.; Fujieda, N.; Sugimoto, H.; Itoh, S. J. Am. Chem. Soc. 2015, 137, 5867– 5870 DOI: 10.1021/jacs.5b01814There is no corresponding record for this reference.
- 5(a) Corona, T.; Pfaff, F. F.; Acua-Pares, F.; Draksharapu, A.; Whiteoak, C. J.; Martin-Diaconescu, V.; Lloret-Fillol, J.; Browne, W. R.; Ray, K.; Company, A. Chem. - Eur. J. 2015, 21, 15029– 15038 DOI: 10.1002/chem.2015018415aReactivity of a Nickel(II) Bis(amidate) Complex with meta-Chloroperbenzoic Acid: Formation of a Potent Oxidizing SpeciesCorona, Teresa; Pfaff, Florian F.; Acuna-Pares, Ferran; Draksharapu, Apparao; Whiteoak, Christopher J.; Martin-Diaconescu, Vlad; Lloret-Fillol, Julio; Browne, Wesley R.; Ray, Kallol; Company, AnnaChemistry - A European Journal (2015), 21 (42), 15029-15038CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)Herein, we report the formation of a highly reactive nickel-oxygen species that has been trapped following reaction of a NiII precursor bearing a macrocyclic bis(amidate) ligand with meta-chloroperbenzoic acid (HmCPBA). This compd. is only detectable at temps. below 250 K and is much more reactive toward org. substrates (i.e., C-H bonds, C-C bonds, and sulfides) than previously reported well-defined nickel-oxygen species. Remarkably, this species is formed by heterolytic O-O bond cleavage of a Ni-HmCPBA precursor, which is concluded from exptl. and computational data. On the basis of spectroscopy and DFT calcns., this reactive species is proposed to be a NiIII-oxyl compd.(b) Pfaff, F. F.; Heims, F.; Kundu, S.; Mebs, S.; Ray, K. Chem. Commun. 2012, 48, 3730– 3732 DOI: 10.1039/c2cc30716bThere is no corresponding record for this reference.
- 6(a) Draksharapu, A.; Codolá, Z.; Gómez, L.; Lloret-Fillol, J.; Browne, W. R.; Costas, M. Inorg. Chem. 2015, 54, 10656– 10666 DOI: 10.1021/acs.inorgchem.5b01463There is no corresponding record for this reference.(b) Pirovano, P.; Farquhar, E. R.; Swart, M.; McDonald, A. R. J. Am. Chem. Soc. 2016, 138, 14362– 14370 DOI: 10.1021/jacs.6b08406There is no corresponding record for this reference.(c) Corona, T.; Draksharapu, A.; Padamati, S. K.; Gamba, I.; Martin-Diaconescu, V.; Acuña-Parés, F.; Browne, W. R.; Company, A. J. Am. Chem. Soc. 2016, 138, 12987– 12996 DOI: 10.1021/jacs.6b07544There is no corresponding record for this reference.
- 7Corona, T.; Company, A. Chem. - Eur. J. 2016, 22, 13422– 13429 DOI: 10.1002/chem.2016024147Spectroscopically Characterized Synthetic Mononuclear Nickel-Oxygen SpeciesCorona, Teresa; Company, AnnaChemistry - A European Journal (2016), 22 (38), 13422-13429CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; iron, copper, and manganese are the predominant metals found in oxygenases that perform efficient and selective hydrocarbon oxidns. and for this reason, a large no. of the corresponding metal-oxygen species has been described. However, in recent years nickel has been found in the active site of enzymes involved in oxidn. processes, in which nickel-dioxygen species are proposed to play a key role. Owing to this biol. relevance and to the existence of different catalytic protocols that involve the use of nickel catalysts in oxidn. reactions, there is a growing interest in the detection and characterization of nickel-oxygen species relevant to these processes. In this Minireview the spectroscopically/structurally characterized synthetic superoxo, peroxo, and oxonickel species that have been reported to date are described. From these studies it becomes clear that nickel is a very promising metal in the field of oxidn. chem. with still unexplored possibilities.
- 8(a) Camasso, N. M.; Sanford, M. S. Science 2015, 347, 1– 7 DOI: 10.1126/science.aaa4526There is no corresponding record for this reference.(b) Riordan, C. G. Science 2015, 347, 1203 DOI: 10.1126/science.aaa7553There is no corresponding record for this reference.
- 9(a) Gray, H. B.; Hare, C. R. Inorg. Chem. 1962, 1, 363– 368 DOI: 10.1021/ic50002a034There is no corresponding record for this reference.(b) O’Halloran, K. P.; Zhao, C.; Ando, N. S.; Schultz, A. J.; Koetzle, T. F.; Piccoli, P. M. B.; Hedman, B.; Hodgson, K. O.; Bobyr, E.; Kirk, M. L.; Knottenbelt, S.; Depperman, E. C.; Stein, B.; Anderson, T. M.; Cao, R.; Geletii, Y. V.; Hardcastle, K. I.; Musaev, D. G.; Neiwert, W. A.; Fang, X.; Morokuma, K.; Wu, S.; Kögerler, P.; Hill, C. L. Inorg. Chem. 2012, 51, 7025– 7031 DOI: 10.1021/ic2008914There is no corresponding record for this reference.
- 10Wieghardt, K.; Schmidt, W.; Herrmann, W.; Kueppers, H. J. Inorg. Chem. 1983, 22, 2953– 2956 DOI: 10.1021/ic00162a037There is no corresponding record for this reference.
- 11Bossek, U.; Nühlen, D.; Bill, E.; Glaser, T.; Krebs, C.; Weyhermüller, T.; Wieghardt, K.; Lengen, M.; Trautwein, A. X. Inorg. Chem. 1997, 36, 2834– 2843 DOI: 10.1021/ic970119hThere is no corresponding record for this reference.
- 12(a) Klamt, A.; Schüürmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 799– 805 DOI: 10.1039/P2993000079912aCOSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradientKlamt, A.; Schueuermann, G.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1993), (5), 799-805CODEN: JCPKBH; ISSN:0300-9580.Starting from the screening in conductors, an algorithm for the accurate calcn. of dielec. screening effects in solvents is presented, which leads to rather simple explicit expressions for the screening energy and its analytic gradient with respect to the solute coordinates. Thus geometry optimization of a solute within a realistic dielec. continuum model becomes practicable for the first time. The algorithm is suited for mol. mechanics as well as for any MO algorithm. The implementation into MOPAC and some example applications are reported.(b) Pye, C. C.; Ziegler, T. Theor. Chem. Acc. 1999, 101, 396– 408 DOI: 10.1007/s00214005045712bAn implementation of the conductor-like screening model of solvation within the Amsterdam density functional packagePye, Cory C.; Ziegler, TomTheoretical Chemistry Accounts (1999), 101 (6), 396-408CODEN: TCACFW; ISSN:1432-881X. (Springer-Verlag)The conductor-like screening model (COSMO) of solvation was implemented in the Amsterdam d. functional program with max. flexibility in mind. Four cavity definitions were incorporated. Several iterative schemes were tested for solving the COSMO equations. The biconjugate gradient method proves to be both robust and memory-conserving. The interaction between the surface charges and the electron d. may be calcd. by integrating over either the fitted or exact d., or by calcg. the mol. potential. A disk-smearing algorithm is applied in the former case to avoid singularities. Several SCF/COSMO coupling schemes were examd. in an attempt to reduce computational effort. A gradient-preserving algorithm for removing outlying charge was implemented. Preliminary optimized radii are given. Applications to the benzene oxide-oxepin valence tautomerization and to glycine conformation are presented.(c) Swart, M.; Rösler, E.; Bickelhaupt, F. M. Eur. J. Inorg. Chem. 2007, 2007, 3646– 3654 DOI: 10.1002/ejic.200700228There is no corresponding record for this reference.
- 13van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597– 4610 DOI: 10.1063/1.46605913Relativistic regular two-component Hamiltoniansvan Lenthe, E.; Baerends, E. J.; Snijders, J. G.Journal of Chemical Physics (1993), 99 (6), 4597-610CODEN: JCPSA6; ISSN:0021-9606.In the present work, potential-dependent transformations were used to transform the four-component Dirac Hamiltonian into relativistic, effective, two-component, regular Hamiltonians. To zeroth order, the expansions give second-order differential equations (just like the Schroedinger equation), which already contain the most important relativistic effects, including spin-orbit coupling. One of the zeroth- order Hamiltonians is identical to the one obtained earlier by Ch. Chang, et al., (1986). By using these Hamiltonians, self-consistent all-electron and frozen-core calcns., as well as first-order perturbation calcns. were done for the uranium atom. They gave very accurate results, esp. for the one-electron energies and electron densities of the valence orbitals.
- 14(a) Swart, M. Chem. Phys. Lett. 2013, 580, 166– 171 DOI: 10.1016/j.cplett.2013.06.045There is no corresponding record for this reference.(b) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098– 3100 DOI: 10.1103/PhysRevA.38.309814bDensity-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.(c) Perdew, J. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33, 8822– 8824 DOI: 10.1103/PhysRevB.33.882214cDensity-functional approximation for the correlation energy of the inhomogeneous electron gasPerdewPhysical review. B, Condensed matter (1986), 33 (12), 8822-8824 ISSN:0163-1829.There is no expanded citation for this reference.
- 16Tak, H.; Lee, H.; Kang, J.; Cho, J. Inorg. Chem. Front. 2016, 3, 157– 163 DOI: 10.1039/C5QI00206K16A high-spin nickel(II) borohydride complex in dehalogenationTak, Hyeonwoo; Lee, Hyunjoo; Kang, Joongoo; Cho, JaeheungInorganic Chemistry Frontiers (2016), 3 (1), 157-163CODEN: ICFNAW; ISSN:2052-1553. (Royal Society of Chemistry)A nickel(II)-borohydride complex bearing a macrocyclic tridentate N-donor ligand, [Ni(Me3-TACN)(BH4)(CH3CN)]+ (Me3-TACN = 1,4,7-trimethyl-1,4,7-triazacyclononane), was prepd., isolated, and characterized by various physicochem. methods, including UV-vis, ESI-MS, IR and X-ray analyses. The structural and spectroscopic characterization clearly shows that the borohydride ligand is bound to the high-spin nickel(II) center in an η2-manner. D. functional theory calcns. provided geometric information of 2, showing that the η2-binding of borohydride to the nickel center is more favorable than the η3-binding mode in CH3CN. The complex is paramagnetic with an effective magnetic moment of 2.9μB consistent with a d8 high-spin system. The reactivity of the high-spin nickel(II)-borohydride complex was examd. in dehalogenation with numerous halocarbons. A kinetic isotope effect value of 1.7 was obsd. in the dehalogenation of CHCl3 by the nickel(II)-borohydride complex. Kinetic studies and isotopic labeling expts. implicate that hydride ion or hydrogen atom transfer from the borohydride group is the rate detg. step. The pos. Hammett ρ value of 1.2, obtained in the reactions of [Ni(Me3-TACN)(BH4)(CH3CN)]+ and para-substituted benzoyl chloride, indicates that the dehalogenation by the nickel(II)-borohydride species occurs via a nucleophilic reaction.
- 17(a) Sun, Y.-K; Kim, M. G.; Kang, S. H.; Amine, K. J. Mater. Chem. 2003, 13, 319– 322 DOI: 10.1039/b209379kThere is no corresponding record for this reference.(b) Gu, W.; Wang, H.; Wang, K. Dalton Trans. 2014, 43, 6406– 6413 DOI: 10.1039/c4dt00308jThere is no corresponding record for this reference.
- 18Rehr, J. J.; Kas, J. J.; Vila, F. D.; Prange, M. P.; Jorissen, K. Phys. Chem. Chem. Phys. 2010, 12, 5503– 5513 DOI: 10.1039/b926434e18Parameter-free calculations of X-ray spectra with FEFF9Rehr, John J.; Kas, Joshua J.; Vila, Fernando D.; Prange, Micah P.; Jorissen, KevinPhysical Chemistry Chemical Physics (2010), 12 (21), 5503-5513CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)A review. The authors briefly review implementation of the real-space Green's function (RSGF) approach for calcns. of x-ray spectra, focusing on recently developed parameter free models for dominant many-body effects. Although the RSGF approach was widely used both for near edge (XANES) and extended (EXAFS) ranges, previous implementations relied on semi-phenomenol. methods, e.g., the plasmon-pole model for the self-energy, the final-state rule for screened core hole effects, and the correlated Debye model for vibrational damping. Here the authors describe how these approxns. can be replaced by efficient ab initio models including a many-pole model of the self-energy, inelastic losses and multiple-electron excitations; a linear response approach for the core hole; and a Lanczos approach for Debye-Waller effects. The authors also discuss the implementation of these models and software improvements within the FEFF9 code, together with a no. of examples.
- 19
In the absence of Ni(II) or 1, NaOCl is stable in CH3CN.
There is no corresponding record for this reference. - 20Inosako, M.; Kunishita, A.; Kubo, M.; Ogura, T.; Sugimoto, T.; Itoh, S. Dalton Trans. 2009, 43, 9410– 9147 DOI: 10.1039/b910237j
This complex is reminiscent of the (μ-η2:η2-disulfido)dinickel(II) complexes with 6-methyl-TPA ligands reported by Itoh et al.; see:
There is no corresponding record for this reference. - 21
The formation of species such as (O═Ni(IV)–O–Ni(IV)═O) is highly unlikely, as it is calculated to lie 45 to 58 kcal mol–1 (depending on functional used, e.g., BP86-D3 and S12g) higher in energy than even the peroxy species 3a. Furthermore, DFT calculations indicate a high tendency for a species to converge to the structure of the 3a species. In addition ligand hydroxylation can be discounted on the basis of the recovery of the Ni(II) complexes, e.g., 1 and 4, once 3 has reacted with solvent or substrate.
There is no corresponding record for this reference. - 22Wieghardt, K.; Bossek, U.; Nuber, B.; Weiss, J.; Bonvoisin, J.; Corbella, M.; Vitols, S. E.; Girerd, J. J. J. Am. Chem. Soc. 1988, 110, 7398– 7411 DOI: 10.1021/ja00230a021There is no corresponding record for this reference.
- 23(a) Hage, R.; Krijnen, B.; Warnaar, J. B.; Hartl, F.; Stufkens, D. J.; Snoeck, T. L. Inorg. Chem. 1995, 34, 4973– 4978 DOI: 10.1021/ic00124a010There is no corresponding record for this reference.(b) Angelone, D.; Abdolahzadeh, S.; de Boer, J. W.; Browne, W. R. Eur. J. Inorg. Chem. 2015, 21, 3532– 3542 DOI: 10.1002/ejic.201500195There is no corresponding record for this reference.
- 24Mandimutsira, B. S.; Yamarik, J. L.; Brunold, T. C.; Gu, W.; Cramer, S. P.; Riordan, C. G. J. Am. Chem. Soc. 2001, 123, 9194– 9195 DOI: 10.1021/ja016209+There is no corresponding record for this reference.
PhTttBu = phenyltris((tert-butylthio)methyl)borate.
- 25Itoh, S.; Bandoh, H.; Nakagawa, M.; Nagatomo, S.; Kitagawa, T.; Karlin, K. D.; Fukuzumi, S. J. Am. Chem. Soc. 2001, 123, 11168– 11178 DOI: 10.1021/ja0104094There is no corresponding record for this reference.
- 26Kieber-Emmons, M. T.; Schenker, R.; Yap, G. P. A.; Brunold, T. C.; Riordan, C. G. Angew. Chem. 2004, 116, 6884– 6886 DOI: 10.1002/ange.200460747There is no corresponding record for this reference.
tmc = 1,4,8,11-tetramethyl-1,4,8,11-tetraazadodecane.
- 27Rettenmeier, C. A.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2015, 54, 4880– 4884 DOI: 10.1002/anie.201500141There is no corresponding record for this reference.
iso-pmbox = bis(oxazolinylmethylidene)pyrrolidine.
Supporting Information
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b04158.
Animated gifs for selected vibrational modes between 517 and 893 cm−1 of 3 (ZIP)
Animated gifs for selected vibrational modes between 536 and 892 cm−1 of 3a (ZIP)
Animated gifs for vibrational modes at 653, 689, 700, 742 cm−1 of 3b (ZIP)
Details of synthesis and charcaterization of complexes 1, 2, and 4, UV–vis absorption, (resonance) Raman, NMR, EPR spectroscopy, ESI-MS, XANES, XES, and computational data (PDF)
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.