You are on page 1of 15

Coordination Chemistry Reviews 257 (2013) 564–578

Contents lists available at SciVerse ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Electronic structure of [Ni(II)S4 ] complexes from S K-edge X-ray absorption


spectroscopy
Matt S. Queen a , Bradley D. Towey a , Kevin A. Murray a , Brad S. Veldkamp b , Harlan J. Byker b ,
Robert K. Szilagyi a,∗
a
Department of Chemistry and Biochemistry, Montana State University, Bozeman, MT 59717, United States
b
Pleotint, LLC, West Olive, MI 49460, United States

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565
2. Experimental techniques and computational methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.1. Preparation of compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.2. Sulfur K-edge X-ray absorption spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.3. XAS data normalization and fitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.4. Electronic structure calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
3. Results and analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
3.1. S K-edge XANES analysis of [Ni(II)S4 ] complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
3.2. S K-edge XANES analysis of S ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
3.3. Development of S 1s → 3p transition dipole integrals for S-ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
3.4. Experimental S 3p contributions to the Ni S bonds in [Ni(II)S4 ] complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
3.5. Theoretical electronic structure of [Ni(II)S4 ] complexes using DFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
4. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577

a r t i c l e i n f o a b s t r a c t

Article history: The nickel ion has a remarkably rich coordination chemistry among the first-row transition
Received 20 March 2012 metals. Complexes with sulfur containing ligands are particularly notable, since they can mani-
Received in revised form 25 July 2012 fest classical/metal-based (innocent) or inverted/ligand-based (non-innocent) behavior depending on
Accepted 28 July 2012
the chemical composition of the S-ligands and the coordination geometry. Using sulfur K-edge X-
Available online 3 August 2012
ray absorption spectroscopy (XAS), we established a spectrochemical series for [Ni(II)S4 ] complexes

Abbreviations: XAS, X-ray absorption spectroscopy; XANES, X-ray absorption near-edge structure analysis (electronic structure); EXAFS, extended X-ray absorption
fine structure analysis (geometric structure); K-edge, X-ray excitation from core 1s level; RIXS, resonant inelastic X-ray scattering; EPR, electron-paramagnetic resonance
spectroscopy; ENDOR, electron-nuclear double resonance, advanced EPR spectroscopy; ESEEM, electron spin echo envelop modulation, advanced EPR spectroscopy; XPS,
X-ray photoelectron spectroscopy; DFT, density functional theory; GGA, generalized gradient approximation for considering 1st derivatives of electron density in DFT cal-
culations; metaGGA, extension of GGA with the 2nd derivatives of electron density; hybrid GGA, electron density functional that mix Hartree–Fock exchange formalism
with GGA exchange functional; hybrid metaGGA, electron density functional that mix Hartree–Fock exchange formalism with metaGGA exchange functional; HOMO, high-
est occupied molecular orbital; LUMO, lowest unoccupied molecular orbital; SUMO, singly unoccupied molecular orbital; Zeff (S), sulfur effective nuclear charge seen by a
given orbital; SPh , aromatic thiolate ligand (actual composition: C6 H4 -Ph-S− ); nbdt, norbornadithiolate ligand; edt, ethylene dithiolate ligand (simplest olefinic enedithi-
olate); dmedt, dimethylethylene dithiolate ligand (olefinic enedithiolate); CNedt, cyanoethylene dithiolate (conjugated olefinic enedithiolate); mnt, maleonitrile dithiolate
ligand (conjugated olefinic enedithiolate); bdt, benzene dithiolate ligand (aromatic enedithiolate); dtc, dithiocarbamate ligand (conjugated); ttctd, tetrathiocyclotetradecane
thiocrownether (aliphatic thioether); NPG, neopentylglycol solvent/ligand; GBL, ␥-butyrolactone solvent/ligand; E0C , energy position for the S 1s → continuum edge jump of
a coordination complex; E0L , energy position for the S 1s → continuum edge jump of a free ligand salt; IC , coordination complex-based, S 1s → 3p transition dipole integral
from independent spectroscopic technique; IL , hypothetical S 1s → 3p transition dipole integral for a ligand not involved in covalent bonding; fwhh, full width at half-height,
line width of XAS features at half amplitude, eV; D0 , normalized intensity or analytical area of an XAS peak, eV.
∗ Corresponding author.
E-mail address: Szilagyi@Montana.EDU (R.K. Szilagyi).

0010-8545/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ccr.2012.07.020
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 565

Keywords: containing thiolate, aliphatic dithiolate, olefinic and aromatic enedithiolate, conjugated dithiocarbamate,
Ni(II) coordination compounds and aliphatic thioether ligands. The pre-edge intensities at the sulfur K-edge follow an increasing trend
X-ray absorption spectroscopy from tetrathiolate through dithiolate and enedithiolate to tetrathioether. In order to obtain quantita-
Ni S bond covalency tive sulfur orbital compositions from XAS data, we generalized the earlier methods of estimating the
Electronic structure
sulfur 1s → 3p transition dipole integral for a broad range of S-ligands by considering chemical shifts
Inverted bonding
Non-innocent ligands
of spectroscopic features due to changes in the sulfur effective nuclear charge and S-ligand coordina-
tion to Ni. The XAS-based experimental orbital compositions are compared with a comprehensive set of
density functional theory-based, electronic structure calculations. The 1,2-dithiolate ligands gave indi-
cation of inverted bonding, independently whether the coordinated sulfur centers are connected by
constrained single or double C,C bonds. Despite intense pre-edge features, both dithiocarbamate and
thioether complexes can be described with classical inorganic bonding description.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction the one-electron oxidation is ligand- and dominantly S-based. The


S K-edge XANES analysis extended the ground state description
Nickel manifests remarkably diverse coordination chemistry of the paramagnetic, formally Ni(III) species, to both its reduced
[1]. In particular, the Ni(II) oxidation state is of interest due to its 3d8 (formally Ni(II)) and oxidized (formally Ni(IV)) forms [13]. For all
valence electron configuration that can exist in paramagnetic S = 1 three members of this electron-transfer series with the same, non-
or diamagnetic S = 0 ground states. For four coordinate complexes, aromatic ligands [19,20], the ground state bonding description can
these spin states correspond with tetrahedral and square planar be described with more than 50% S contribution per electron hole
coordination geometry, respectively. Notably, the paramagnetic to the redox active, unoccupied frontier orbital. Due to the diamag-
tetrahedral state manifests Jahn–Teller distortion, which renders netic ground state of the d8 Ni(II) complex with a single empty
the symmetry of the [Ni(II)S4 ] moiety to be effectively D2d . Further- d-orbital, and thus the presence of spin up (␣) and spin down (␤)
more, the large steric bulk of the substituted S-ligand contributes electron holes, this corresponds to more than 1 electron total S con-
to the elongated Ni S bond lengths and may lower the symmetry tribution for both spin orbitals. This also translates to more than
to S4 [2]. For the diamagnetic complexes, the effective symmetry 25% Ni S bond covalency or more than 0.25 e− donation per each
is D2h due to the coordination environment. The spectrochemical enedithiolate ligand per electron hole, which represents close to
series of weak ␴- and/or ␲-donor ligands versus the strong ␴-donor doubling of covalent bonding relative to the tetrahedral tetrathio-
and ␲-acceptor ligands predetermines the central Ni(II) ion’s coor- late complex [3].
dination geometry and spin state. This is less straightforward for Due to the metal-based, innocent electronic structure, the effec-
the family of S-containing ligands with intermediate ligand field tive and formal oxidation states of the Ni center in a [Ni(II)(SR)4 ]2−
strength. complex are practically identical. However, this is not true for the
In this comparative study, we focused on homoleptic [Ni(II)S4 ] non-innocent, enedithiolate Ni complex ([Ni(dmedt)2 ]2− ), where
coordination compounds that span both spin states, both coor- the effective oxidation state is considerably reduced. This remains
dination geometries, anionic, neutral, and cationic complexes unchanged throughout the 2-electron electron-transfer series for
(Scheme 1). The electronic structure of the high spin, tetrahedral [Ni(dmedt)2 ]− and [Ni(dmedt)2 ]. The non-innocent bonding sce-
complex [Ni(II)(SPh )4 ]2− of this series with aromatic thiolate lig- nario and the corresponding reduction of the Ni effective oxidation
ands (S-Ph = S-2-Ph-C6 H4 ) has already been discussed in detail state versus its formal Ni(II) state can be rationalized by molecu-
[2,3]. This complex can be considered as a representative exam- lar orbital theory that presents an inverted orbital ordering with
ple for the classical inorganic or normal bonding scenario with the S donor orbitals being at higher energy than the vacant Ni
unoccupied, degenerate metal d-orbitals (3dxz and 3dyz of the t2 - 3d orbital. The efficient ␴- and reasonable ␲-overlap in square
set) lying above the symmetry adapted combination of donor S 3p planar complexes between the S lone pairs and the vacant Ni
orbitals in molecular orbital theory or S lone pairs in a valence bond 3d orbitals facilitates intramolecular electron-transfer and con-
picture. Moreover, the sulfur K-edge X-ray Absorption Near-Edge tributes to the reduction of the formally Ni(II) to effectively Ni(0)
Spectroscopic (S K-edge XANES) analysis of the [Ni(II)(SPh )4 ]2− state with concomitant oxidation of the enedithiolate ligands
defined about 33% S 3p character in each of the partially occu- from their thiolate form toward their radical anionic and then
pied Ni 3d-orbitals [3]. This corresponds to approximately 17% further on to their thione form. In this paper, we extend the
Ni S bond covalency or a donation of 0.085 electron (e− ) per each previously studied [Ni(II)S4 ] complexes with three new mem-
ligand from the S lone pairs to each of the electron holes. Another bers: the bis(1,2-norbornadithiolato)nickel(II) ([Ni(nbdt)2 ]2− ),
representative example for the remarkable coordination chem- bis(diethyldithiocarbamato) nickel(II) ([Ni(dtc)2 ]), and the tetrathi-
istry of S-based ligands is the family of bis(enedithiolate)nickel acyclotetradecane coordinated nickel(II) ([Ni(ttctd)]2+ ) complexes.
complexes [4–8] that are also known as bis(dithiolene)nickel com- These complete a spectrochemical series for the [Ni(II)S4 ] motif
plexes. The formal Ni oxidation states of these complex range (Scheme 1) with constrained aliphatic dithiolate, conjugated
from II to IV in [Ni(dmedt)2 ]2− to [Ni(dmedt)2 ], respectively. dithiocarbamate, and aliphatic tetrathioether complexes.
These square planar complexes manifest ligand-based reactivity The existence of the latter complex is notable, since only
[9–11] and inverted bonding [12,13] that have been collectively a few thioether ligands coordinate to Ni(II) [21]. In addition
referred to in the literature as non-innocent coordination behavior to its compositional peculiarity, the [Ni(ttctd)]2+ complex man-
[14–16]. Advanced electron paramagnetic spectroscopic tech- ifests an ideal behavior in a reversible, ligand exchange-based
niques [7,17,18] indicated the non-innocent nature of the formally thermochromic process [22]. The ␥-butyrolactone (GBL) solu-
bis(enedithiolate)nickel(III) complex. tion of the [Ni(ttctd)](ClO4 )2 salt with 30-fold neopentyl glycol
The quantitative analysis of the hyperfine coupling constants (NPG) manifests a reversible color change for the tempera-
defined about 80% spin localization at the central [NiS4 ] unit ture range of 25–100 ◦ C. The solution is almost colorless at
with only about 20% localized on the metal. The inverted con- low temperature, where the Ni2+ ions are coordinated with
tributions from the metal and the ligand to the singly occupied the NPG ligands. As the temperature increases the [Ni(NPG)3 ]2+
␲*-orbital indicate that in going from a formally Ni(II) complex complex converts into [Ni(ttctd)]2+ with max = 503 nm and
566 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

Scheme 1.

ε ≈ 370 dm3 mol−1 cm−1 in GBL (red color). The estimated of the given study is to provide a comprehensive understanding
thermodynamic parameters for the thermochromic process are of metal–ligand bonding in Ni(II) coordination complexes with
ideal due to the large enthalpy (H = 60 kJ mol−1 ) and large entropy thermochromic behavior by X-ray absorption spectroscopy and
(S = 190 J mol−1 K−1 ) changes [23]. The general equilibrium equa- correlated electronic structure calculations. In addition, the exem-
tion for ligand-exchange thermochromic reaction involving Ni(II) plary thermochromic behavior of the Ni-thioether complex also
is as follows: serves as an important control for evaluating the Ni S(thioether)
bond covalency.
Ni(II)(low ε ligand)n + m(high ε ligand)  Ni(II)(high ε ligand)m The X-ray absorption spectroscopy (XAS) probes core-level exci-
+ n(low ε ligand) (1) tations, yet it can be used to obtain quantitative ground state
information [24]. The analysis of the spectroscopic features below
where the low and high ε ligands are defined depending on whether the ionization threshold of the S 1s orbitals (∼2472 eV) can provide
they contribute to the colorless and colored coordination complex, information about the amount of S 3p character in the low lying,
respectively. These ligands for the above thermochromic system unoccupied frontier molecular orbitals. The essential information
are the diol solvent (NPG) and thiocrownether (ttctd) molecules needed to obtain experimental orbital compositions from exper-
with n = 3 and m = 1 stoichiometry giving octahedral and square pla- iment is the knowledge of a S 1s → 3p transition dipole integral
nar coordination complexes, respectively. One of the motivations (I). This requires the availability of independent spectroscopic
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 567

information (EPR/ENDOR/ESEEM or XPS for example) about the optimal saturated basis set for studying the electronic structures of
electronic structure of a given complex. The relevant spectroscopic Ni containing metalloenzymes, and organometallic complexes, as
features in XAS are the well-resolved (pre-edge) and overlapping well as for designing and optimizing Ni(II)-based thermochromic
(rising-edge) features with the ionization threshold (edge jump). systems.
Using a dipole expression [25] as shown below, the area or inten-
sity (D0 , eV/absorber) under a given pre-edge or rising-edge feature 2. Experimental techniques and computational methods
can be correlated with the amount of S 3 p character per electron
hole (˛2 , electron hole covalency): 2.1. Preparation of compounds
1 h 2
D0 = ˛ I(S 1s → 3p) (2) The Na2 S salt [50] and Ni(dtc)2 complex [51] were pre-
3N
pared and characterized according to literature procedures. The
where h represents the number of electron holes in the probed Li2 nbdt salt and the corresponding (Ph4 P)2 Ni(nbdt)2 complex were
unoccupied orbitals, N the number of absorbers, the 1/3 factor is provided by the late Prof. Michelle Millar, State University of
from the x, y, z dependence of the transition dipole expression [26], New York, Stony Brook [52]. The tetrathiacyclotetradecane (ttctd)
and the I(S 1s → 3p) is the transition dipole integral (eV) as proba- thiocrownether was purchased from Sigma–Aldrich and used with-
bility for an excitation of an electron from the sulfur 1s core orbital out further purification and characterization. An ∼5 mM solution
to the valence S 3p level. Although the core electron dipole integral of [Ni(ttctd)]2+ complex was freshly prepared before the XAS mea-
can be estimated from wave function- or density functional-based surements with a slight excess of Ni(ClO4 )2 salt in freshly distilled
calculations [27], for practical use one needs to consider experi- ␥-butyrolactone (GBL). The solid sample was isolated as described
mental factors, such as beamline components, detection method, in Ref. [21]. The XAS spectra of the solution and solid samples were
sample preparation protocols, etc. [28]. found to be practically identical with negligible distortion in peak
A S 1s → 3p transition dipole moment was first determined for amplitudes due to self-absorption effects. All samples measured
thiolate ligands coordinated to a Cu(II) ion [29]. This was later by us were characterized by comparing their solution and/or solid
extended first to sulfide [30] and then to the enedithiolate [13] state UV/vis spectra to literature values [21,51,52].
ligands. A different dipole moment is needed for each S-ligand,
since the sulfur 1s orbital energy is inherently linked to the effec- 2.2. Sulfur K-edge X-ray absorption spectroscopy
tive nuclear charge (Zeff (S)) and thus the S atomic charge. An earlier
study [31] established a linear relationship between the S atomic The S K-edge data for [Ni(SPh )4 ]2− [3], [Ni(dmedt)2 ]2− [13],
charges and the most intense XANES features corresponding to [Ni(bdt)2 ]2− [53], and [Ni(mnt)2 ]2− [33] were taken from literature.
excitations into S C ␴* or S 4p level. This observation allowed for All the other complexes and the corresponding free ligands were
the extension of known dipole integrals to other S-ligands. Overall, either measured at beamline 6-2 ((Ph4 P)2 Ni(nbdt)2 and Li2 (nbdt),
there are two conceptually different approaches in the literature for (ClO4 )2 Ni(ttctd) and ttctd) or 4-3 (Ni(dtc)2 and Nadtc) of the Stan-
obtaining experimental S 1s → 3p transition dipole integrals for a ford Synchrotron Radiation Lightsource. Repeated measurements
transition metal complex with vacancy in the d-manifold. The first of the samples did not reveal any significant variation among data
method [29,30,32] used the XAS spectra of S-ligands coordinated sets from the different beamlines for solid samples. All XAS mea-
to first row transition metals and independent spectroscopic infor- surements were carried out at room temperature for sample cells
mation to assign pre-edge intensities to orbital compositions. The that were prepared in nitrogen or argon gas filled glovebox. While
second method compared spectroscopic features of free and coor- mounting the samples at the beamline, they were protected with
dinated ligands and employed a linear relationship between the S a few ␮m thick Mylar window to eliminate sample degradation
charge as represented by the S 1s → 4p excitation energy positions due to oxidation or moisture exposure. Sample decomposition was
and the transition dipole integral [13,33]. The given comprehensive clearly identified by the appearance of the intense sulfate feature
data for Ni(II) coordinated S-ligands provided the opportunity for at around 2482.5 eV [54]. Spectra with extended energy ranges in
merging the two approaches and defining a general method for esti- Figure S1 clearly indicate the lack of sulfate in all data considered.
mating transition dipole integrals using only XAS data of [Ni(II)S4 ] The solid samples were ground with boron nitride and pasted onto
complexes and the corresponding free S-ligands. Kapton tape. Any excess sample was scraped off from the tape to
An additional implication of our study is the use of experi- minimize the distortion of peak intensities due to self absorption
mental electronic structure description to evaluate the accuracy effects [55]. The beamlines were optimized to maximize incident
of electronic structure calculations. For Cu(II) complexes [34], it beam intensity at the end of the scan (2750 eV). All data presented
has been shown that a hybrid functional with approximately 38% here were collected with fluorescence detection either using a
Hartree–Fock exchange mixed with Becke 1988 density functional nitrogen gas purged, multi-element Lytle fluorescence emission
exchange [35] reproduced the experimental orbital composition detector (EXAFS Co) or a passivated implanted planar silicon detec-
of the ground state of [CuCl4 ]2− [36,37]. The same hybrid func- tor (PIPS, CANBERRA). Importantly, due to the observed intensity
tional also provides reasonable electronic structure description of differences between electron yield and fluorescence emission data,
a series of Cu-containing metalloproteins [38–42]. It is important the pre-edge intensity analysis and the transition dipole integrals
to add here that the appropriate treatment of protein-, crystal-, presented here are limited to reasonably thin samples with the
or solvent-environmental effects in computational modeling is as latter detection technique.
critical as the selection of the most reasonable density functional
[41–45]. Follow up studies investigated the performance of various 2.3. XAS data normalization and fitting
density functionals for a group of Fe-S clusters and found preference
for another hybrid functional with less Hartree–Fock exchange (5%) All data was background subtracted and normalized using the
[46,47]. This functional was essential in predicting the most rea- Automated Data Reduction Protocol developed by Gardenghi [56].
sonable composition of the elusive light atoms in Mo-nitrogenase The normalized data were fit using Peak Fit v4.12 (SeaSolve).
[48] and FeFe-hydrogenase [49] a priori to experimental confir- The details of all fits are shown in the supporting information
mations. Given the success of these empirically developed hybrid (Figure S2–S9). Due to the comprehensive set of spectra for both the
functionals, we wish to extend this approach to Ni(II) complexes Ni(II) complexes and the free ligand salts, we were able to employ
and thus provide a spectroscopically calibrated method with an a common protocol. All resolved features were fit by pseudo-Voigt
568 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

lines for non-degenerate excitations below the most intense, S C Similarly to the experimental covalency from XAS as defined
␴* features including the S C ␲*, and even Li S ␴* orbitals. Transi- by Eq. (2), the S 3p character of LUMO or S 3p contributions to
tion envelopes were fit using the Gaussian + Lorentzian Amplitude covalent bonding per electron hole (or per spin orbital) was
Function in PeakFit with unrestricted Gaussian/Lorentzian mix- employed. This is most consistent with a molecular orbital theory-
ing ratio, while single or degenerate double electron hole-based based bonding analysis. The per-hole covalency already takes into
excitations were fit with a Gaussian/Lorentzian ratio of 1:1. The cor- account all the contributions from all symmetry allowed S donors.
responding spectroscopic features were subtracted from the free For the diamagnetic complexes, the orbital covalency is simply the
ligand salt spectra to obtain a reasonable rising-edge background double of per-hole covalency. Furthermore, multiplication of the
for the complexes. Due to the covalent Li S bonding in Li2 nbdt ‘per-orbital’ values with the number of unoccupied orbitals defines
salt, the reference as “free ligand” spectrum is not strictly correct; the total sulfur covalency. Since the Ni(II) complexes have more
however, by removing the rising-edge feature we can obtain a rea- than one vacant d-electron hole, the upper limit of covalency is
sonable estimate for the free ligand spectrum (Figure S3). A free 2 e− . Division of the total S covalency by the number of bonds or
ligand spectrum was fit to create a user defined function, which number of atoms gives the per-bond or per-absorber covalency val-
was then used to subtract C S ␴* transition-based rising-edge and ues that are mainly used in valence-bond theory-based discussions
edge-jump features by using the second derivative spectra to align of metal–ligand bonding.
the most intense transitions. The remaining pre-edge (Ni S bond-
ing) and rising-edge (S C ␲* bonding) peaks after the subtraction of
the free ligand spectrum were fit with Gaussian/Lorentzian ratio of 3. Results and analysis
1:1 and a linked linewidth of about 1.2 eV. In order to avoid unrea-
sonable distortion and large variations in intensities for ill-resolved 3.1. S K-edge XANES analysis of [Ni(II)S4 ] complexes
features, additional fits were obtained in consecutive steps by grad-
ually allowing for the linewidths, amplitudes, and lastly the energy The comprehensive list of Ni(II) complexes used in the given
positions to be altered. The final fits were obtained by relaxing all study includes a paramagnetic (S = 1) tetrahedral [Ni(SPh )4 ]2−
fit parameters except the Gaussian/Lorentzian mixing ratio. complex, and a series of diamagnetic (S = 0) square planar
dithiolate [Ni(nbdt)2 ]2− , aliphatic enedithiolate [Ni(dmedt)2 ]2− ,
aromatic enedithiolate [Ni(bdt)2 ]2− , dithiocarbamate [Ni(dtc)2 ],
2.4. Electronic structure calculations and tetrathiocrownether [Ni(ttcd)]2+ complexes. Fig. 1 compares
the S K-edge pre-edge and rising-edge features for the above
All electronic structure calculations were performed using listed complexes. The straightforward molecular orbital picture
Gaussian 09 [57] in the presence of a polarizable dielectric (Scheme 2) of these complexes allows for the assignment of the
environment [58–63] solvent parameters for acetonitrile with a resolved spectroscopic features below and along the rising-edge.
dielectric constant of 35.6 and solvent radius of 1.81 Å. The com- The lowest energy pre-edge feature can be associated with the
putational models were created from crystallographic structures excitation of the S 1s core electron into the unoccupied, Ni S
([Ni(II)(SPh )4 ]2− [2], [Ni(II)(nbdt)2 ]2− [52], [Ni(II)(dmedt)2 ]2− antibonding ␲* or ␴* orbitals for the paramagnetic or diamag-
[19], [Ni(II)(bdt)2 ]2− [64], [Ni(II)(CNedt)2 ]2− [65], [Ni(II)(mnt)2 ]2− netic complexes, respectively. For the [Ni(dmedt)2 ]2− , [Ni(bdt)2 ]2− ,
[66,67], [Ni(II)(dtc)2 ] [68], [Ni(II)(ttctd)]2+ [21]) and used without [Ni(dtc)2 ] complexes, the next feature corresponds to the excita-
structural optimizations. A summary of intramolecular distances tions into C S ␲* orbitals, followed by the C S ␴* transitions as
and bond angles relevant to the coordination geometry is given in most intense spectroscopic lines (white lines) for all complexes. The
Table S1. In order to evaluate the uncertainties in LUMO compo- resolution of the S K-edge XAS spectra (∼1.2 eV linewidth for single
sition due to the ill-defined crystallographic H atomic positions, electron excitation) limits unique assignments of any higher lying
the structures of representative examples of the above series were excitations, but from frontier orbital energy levels it is expected
partially optimized. It was determined that changes in the S con- that the S 4p-based bound states will follow with the overlapping
tributions to the LUMO were less than 0.05 e− per electron hole. A edge-jump that corresponds to the ionization threshold of the S 1s
set of representative exchange and correlation density functionals orbital into the continuum. The approximate position of the edge
was chosen according to the rungs of Perdew’s ladder of functionals jump thus can be estimated from the first inflection point after the
[69,70]. From the functionals with generalized gradient approxima- C S ␴*-based transitions or the ‘white line’ feature. Fig. 1B shows
tions rung the commonly used BLYP [35,71], BP86 [35,72], mPW the first derivative spectra, which can be used to obtain the energy
[73], OLYP [71,74,75], PBE [76], and PW91 [77,78] functionals were positions of inflection points. In general, it is common to correlate
employed. The M06L [79], TPSS [80], and TPSSLYP1W [81] function- the rising-edge inflection points as determined by the first max-
als were selected for representing the metaGGA rung. As a bridge imum of the first derivative along the rising-edge energy region
between the two rungs, we used a set of hybrid functionals, that with the absorber’s effective nuclear charge [102]. However, this is
mix Hartree–Fock exchange and thus ionic character into either only valid for complexes with the same overall electronic structure.
GGA (B1LYP [71,82], B3LYP [71,83], BHandHLYP [71,84], MPW1K In practice, the rising-edge inflection points at the ligand K-edge
[85], mPW1PBE [73,76], MPW3LYP [86], MPW1LYP [87], and O3LYP are only useful for monoatomic/ionic ligands, such as chloride [26].
[71,88]) or metaGGA (M06 [89], M062X [89], M06HF [90,91], and The comparison of the rising-edge inflection point positions of
TPSSh [80]) exchange functionals. We employed the def2-TZVP an aliphatic thiolate coordinated complex to olefinic or aromatic
basis set [92] from the EMSL Gaussian Basis Set Exchange [93,94] enedithiolate complexes will result in erroneous effective nuclear
that can be considered saturated for electronic structure and thus charge estimates due to the rising-edge originating from a C S ␴*
orbital compositions. The theoretical total S composition of the or C S ␲* based orbitals, respectively. Fig. 1C presents the second
unoccupied frontier orbitals corresponding to the Ni S bonds were derivative spectra that can be used to obtain information about the
obtained from Bader’s ‘atoms in molecules’ population analysis number of peaks and their energy positions. These are important in
[95–98] using AIMAll program [99] in addition to the conventional quantitative analysis as the minima can guide the fitting procedure
Mulliken population analysis (MPA) [100] and Weinhold’s Natu- and effectively reduce the number of parameters to be fit.
ral population analysis (NPA) [101] as implemented in Gaussian09. From Fig. 1C, the pre-edge energy position gradually shifts
The latter two orbital-based methods provided the experimentally in going from tetrathiolate, bis(dithiolate), bis(enedithiolate),
directly comparable S 3p characters. bis(dithiocarbamate), and tetrathiocrownether coordinated Ni(II)
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 569

Fig. 1. Comparison of S K-edge XANES spectra of Ni(II) coordinated S-ligands considered in this work (nbdt = 1-2-norbornadithiolate, edt = ethylenedithiolate,
bdt = benzenedithiolate, mnt = maleonitriledithiolate, dtc = diethyldithiocarbamate, ttctd = tetrathiacyclotetradecane) (A) normalized data, (B) first derivative spectra with
inflection points in eV, (C) second derivative spectra with peak positions in eV. The S K-edge data for [Ni(SPh )4 ]2− [3], [Ni(dmedt)2 ]2− [13], [Ni(bdt)2 ]2− [53], and [Ni(mnt)2 ]2−
[33] were taken from literature.

complexes. However, this energy position is affected by both ligand–ligand repulsive interaction raises the energy of the S 1s
the sulfur effective nuclear charge (ligand-based chemical shift) orbital with respect to those in the tetrahedral Ni S(thiolate) com-
and the Ni d-manifold energy (coordination-based chemical shift). plex. The effect of the ligand-based chemical shift is most drastic for
In going from the tetrathiolate to the dithiolate complex, the the thioether complex with the highest Zeff (S) among all [Ni(II)S4 ]
dominant change is the change in coordination geometry from complexes. This results in a shift to higher energy by approximately
tetrahedral to square planar. Therefore, there is positive energy 1 eV of the pre-edge feature and the edge jump relative to the
shift in the Ni d-manifold, since the better overlap in the square pla- dithiocarbamate complex.
nar geometry creates a larger ligand field splitting and thus pushes In order to carry out the comparison of the pre-edge features,
the frontier 3d orbitals to higher energy. In addition, the better we have chosen to eliminate the rising-edge features using the
overlap allows for greater mixing of the S-based orbitals of the corresponding uncoordinated ligand salt spectra. Figures S2–S10
ligand and the vacant metal orbitals (see Section 4 for a quantita- show the data points for the coordination complex, free lig-
tive comparison of the overlap integrals). The greater mixing of S 3p ands (solid diamonds and hollow triangles, respectively), the
with Ni 3d orbitals is well demonstrated by considerable increase shifted/scaled free ligand spectra (thin solid line), the deconvo-
in pre-edge intensity in [Ni(nbdt)2 ]2− relative to the [Ni(SPh )4 ]2− luted pre-edge feature (shaded pseudo-Voigt feature with 50%/50%
complex. However, the reduced distance of the two S atoms in mixed Gaussian/Lorentzian line shape), and the sum of the pre-
the bis(dithiolate) complex due to the rigid C C bond within a edge feature and shifted free ligand spectra (thick line). Fig. 2
constrained cyclic aliphatic framework or the shorter C C double compares the pre-edge features that correspond only to the
bond within a olefinic or aromatic enedithiolate and the smaller S 1s → ␸(Ni 3d S 3p) ␲* and/or ␴* excitation. As defined by
S· · ·Ni· · ·S bite angle induces a considerable S ↔ S repulsion. This the above dipole integral expression, the intensities of these

Scheme 2.
570 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

The pre-edge peak intensities already indicate the amount of


S → Ni charge donation as a result of covalent bond formation
between the symmetry adapted linear combination of S 3p orbitals
and the Ni 3d orbital(s). It is important to note that the peak
areas are given as exact analytical areas and not as estimated val-
ues from the product of peak amplitude and full width, which is
approximately 27% lower. The significant deviation in Ni S bond-
ing between the tetrahedral and square planar complexes is well
represented by the large differences in pre-edge peak intensities.
Tying the two thiolates together with a C C bond in a constrained
aliphatic ligand frame drastically increases the pre-edge inten-
sity in [Ni(nbdt)2 ]2− (1.28 eV) relative to [Ni(SPh )4 ]2− (0.56 eV).
Replacing the C C link between the two adjacent sulfur donors
in aliphatic dithiolates with a shorter and more constrained C C
link in enedithiolates, further increases the pre-edge intensity to
the highest value of 1.84 eV for [Ni(dmedt)2 ]2− among the listed
Fig. 2. Comparison of S K-edge pre-edge features after subtraction of rising-edge complexes in Table 1. As the peripheral methyl groups of the
features using the free ligand salt spectra (Figures S2–S9). [Ni(dmedt)2 ]2− are stepwise replaced by electron withdrawing
cyanide substituents, the pre-edge intensity drops due to delocal-
spectroscopic features correspond with the amount of S 3p mix- ization of the S 3p lone pairs into the olefinic conjugated system.
ing with the vacant Ni 3d orbital. This is commonly referred to as Remarkably, this delocalization is also present for the aromatic
the Ni S bond covalency or S contribution to bonding. In order to dithiolate, but to a lesser extent as indicated by the modest change
quantify this information we need to determine a transition dipole in the pre-edge intensity (0.86 eV) in the [Ni(bdt)2 ]2− complex.
moment for each of the complexes. As we approach the higher energy pre-edge features such those
Table 1 summarizes the energy positions, amplitudes, in Ni(dtc)2 and [Ni(ttctd)]2+ there are some notable peculiarities.
linewidths, and peak intensities that were obtained from fit- For the former complex, we still see an intense pre-edge feature
ting the rising-edge corrected, pre-edge features from Fig. 2 with (1.60 eV) despite the non-ideal S Ni S bite angle (79◦ ) for S lone
a pseudo-Voigt line. The pre-edge energy positions in Table 1 pair and Ni 3dxy overlap and the reduction of the formal oxidation
already indicate the presence of different S-ligand environment state of the S absorbers. In going from the dtc− to the ttctd ligand,
and thus the different nature of the Ni S bonding. The lowest the pre-edge intensity remains practically the same (1.59 eV), but
energy position observed for the pre-edge was measured for the the S effective nuclear charge is further reduced to formally a neu-
tetrathiolate complex due to the least covalent Ni S bonding tral S state. The thioether is a weak donor ligand; however, in the
(smallest D0 values) [3] that corresponds to the most oxidized given ligand arrangements, the chelating angles (89.7◦ ) are ideal for
Ni(II) with the smallest ligand field splitting due to the tetrahedral efficient Ni SR2 overlap and therefore S → Ni electron density shift
coordination geometry. The pre-edge features for the dithiolates results in covalent Ni S bonds. The reversible, ligand-exchange
and enedithiolates can be found in the 2470.9–2471.3 eV energy thermochromic behavior of the [Ni(ttctd)]2+ complex already hints
range, which is about 0.6 eV shifted up in energy relative to that the covalency will be considerably reduced relative to the
the tetrathiolate complex. However, this modest shift actually enedithiolate complexes, which do not show any thermochromic
corresponds to a large reduction of the effective nuclear charge behavior. Overall, the above trends clearly indicate that the pre-
of the Ni(II) due to the considerable S → Ni electron donation as edge intensity itself cannot be the measure of bond covalency when
well as the increase in the ligand field splitting due to the square compared for various S-ligands.
planar coordination environment. The large change is masked by In order to obtain S 3p orbital compositions of the Ni S anti-
the lower energy position of the edge jump for the square planar bonding orbital from pre-edge intensities at the S K-edge, we need
complexes by about 1 eV as can be seen in the last column of to determine the transition dipole integral or dipole moment for
Table 1. There is another considerable stepwise jump in both the the S 1s → 3p excitation for all ligands and their complexes. From
pre-edge and the edge jump positions for the dithiocarbamate and complementarity, this S 3p contribution reflects the S character of
the thioether ligand containing complexes. the Ni S bonding orbital. This dipole integral, I (eV), can then be
used to convert the normalized intensity (D0 , eV/absorber) to total
intensity (D0 × N, where N is 4 as the number of S absorber for the
Table 1 series of complexes from Table 1), and then to total S 3p character
Fitting results for the pre-edge features in Fig. 2 and related energy positions from
Fig. 1.
(average S 3p character or covalency × number of electron holes,
˛2 × h) as defined by Eq. (2). Note that the energy dimension of
Complexes E, eVa A, –b fwhh, eVc D0 , eVd E0C , eVe the intensity (D0 ) and therefore the dipole integral is due to the
 2− f integration of peak area defined by a linewidth in eV and dimen-
[Ni(SPh )4 ] 2470.8 0.40 1.10 0.56 2475.1
[Ni(nbdt)2 ]2− 2471.1 0.95 1.07 1.28 2474.1 sionless normalized peak amplitude. The total S 3p character can
[Ni(dmedt)2 ]2− 2471.3 1.13 1.28 1.84 2474.8
then be expressed as per-hole or per-bond covalency. The per-hole
[Ni(bdt)2 ]2− 2471.4 0.60 1.13 0.86 2474.8
[Ni(CNedt)2 ]2− 2471.2 1.15 0.94 1.37 2474.6 value has a theoretical maximum of 100%, which in reality cannot
[Ni(mnt)2 ]2− 2471.3 1.15 0.97 1.40 2474.3 be more than about 80% for the given set of S ligands, since each
[Ni(dtc)2 ] 2471.5 1.28 0.99 1.60 2474.9 ligand donor orbital may have non-negligible C and H contribution
[Ni(ttctd)]2+ 2472.6 1.29 0.97 1.59 2475.8 in addition to the Ni character of a given LUMO. Even in the case of
a
Pre-edge peak position. an inverted bonding scheme as discussed in literature for enedithi-
b
Pre-edge peak intensity. olate ligands [12,13] there will be about 20% Ni contribution to
c
Half-max full width.
d
the LUMO. As discussed above, not all complexes in Table 1 with
Normalized, analytical pre-edge peak area.
e
Estimated energy position of the edge inflection point. large pre-edge intensities (D0 ) are expected to have inverted bond-
f
Average values are given from the three difference fits shown in Figure S2 that ing pictures, and thus the S contributions will be below 50% as a
match the published pre-edge intensity in Ref. [3]. result of the composition and structure of each S-ligand. In order to
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 571

Fig. 3. Comparison of S K-edge XANES spectra of S-ligands considered in this work. (nbdt = 1-2-norbornadithiolate, edt = ethylenedithiolate, bdt = benzenedithiolate,
mnt = maleonitriledithiolate, dtc = diethyldithiocarbamate, ttctd = tetrathiacyclotetradecane) (A) normalized data, (B) first derivative spectra with inflection points in eV,
(C) second derivative spectra with peak positions in eV. The S K-edge data for edt2− [13] were taken from literature.

evaluate the difference between the S-ligands within the complex Fig. 3C compares the second derivative spectra that can be
and their free, ionic forms we considered the S K-edge XAS spectra used to estimate the peak positions of each rising-edge feature if
of a series of S-ligands (Fig. 3). resolved. It is notable that with the exception of sulfide the fea-
ture corresponding to S 4p-based excitations can only be assigned
ambiguously. Therefore, we define the energy position of the first
3.2. S K-edge XANES analysis of S ligands inflection point after the last resolved C S ␴* feature to be the
experimental measure of the Zeff (S) seen by the S 1s core electron.
We can separate the effects of ligand-based and coordination- From a practical peak fitting perspective, this is generally the lower
based chemical shifts on the energy positions and intensities of limit of an energy position where a step function is introduced that
spectroscopic features in Figs. 1 and 2 by comparing the S K-edge represents the ionization threshold or the edge jump. By mapping
spectra of S-ligands in Fig. 3. The comprehensive set of ligands cover out the effect of Zeff (S) on the S K-edge XANES features, we are in the
all three important sulfur oxidation states from formally S2− in sul- position to define the influence of the coordination-based chemical
fide to S0 in thioether. The enedithiolates can be considered to have shift for the comprehensive series of Ni(II) complexes.
a formally S1− charge, however due to conjugation the effective
charge is less negative. Similarly, the formal sulfur charge in the 3.3. Development of S 1s → 3p transition dipole integrals for
dithiocarbamate ion is S−0.5 , which is effectively closer to a neu- S-ligands
tral S(thioether) than to the anionic S(thiolate) due to the extensive
delocalization of the out-of-plane ␲-system involving the lone-pair The first S 1s → 3p transition dipole integral was developed for
of the amine N. These qualitative considerations on the basis of the Cu(II) ion coordinated electron-transfer site of plastocyanin and
valence bond theory are well reproduced by the changes in spec- a related model complex [29]. This was further confirmed by a
troscopic features as shown in Fig. 3. With the exception of the two very detailed and comprehensive spectroscopic evaluation of three
Li+ salts, all complexes can be considered as in their free ionic form variants of the Cu(II) site in stellacyanin [32]. The analytical peak
with minimal covalent interactions. Thus, trends in Fig. 3 describe intensity of 1.54 eV in Table 2 was calculated from the approxi-
the ligand-based chemical shifts. It is relevant to the dipole integral mated peak intensity of 1.21 eV and applying a 1.27 conversion
discussion and important to point out that consideration of rising- factor as a difference between the analytical integrated areas and
edge features (first maximum of the first derivative spectra along the amplitude × full linewidth approximation. This pre-edge inten-
the rising-edge) as the measure of Zeff (S) does not give an chemi- sity was associated with about 45 ± 3% S 3p character. This agrees
cally acceptable trend. This is due to the different origin (Li S ␴*, well with complementary Cu hyperfine coupling constant analy-
S C ␴*, S C ␲*, S 4p) of the rising-edge features. The white lines sis from EPR and Cu L-edge spectroscopic features [103]. Notably,
of all S-ligands can be related to the excitation into C S ␴*-based since the spectra were obtained for a metalloprotein with other S
orbital except for the sulfide ligand. However, the origin of the spec- absorbers that do not contribute to the pre-edge feature, therefore
troscopic features at the lower energy side of the white lines varies the edge inflection point could not be determined.
depending on the presence of C S ␲* orbitals (in enedithiolates and The transition dipole integral for a more reduced form of sulfur
dithiocarbamate) or covalent interactions between the counter ion was obtained from analysis of the pre-edge feature of the CsFe(III)S2
and the sulfur (in Li2 nbdt and Li2 bdt salts). salt with bridging sulfides. The analytical intensity of 3.12 eV
572 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

Table 2
Comparison of experimental data (XPS or EPR) defining S 1s → 3p transition dipole integral (IC ) for S-ligands bound to transition metal ions.

Ligand Complex D0 , eVa Nb hc ˛2 , % IC , eV E0C , eVd E0L , eVd



Bridging sulfide [Fe(III)S2 ] 3.12 2 5 42 8.90 2473.6 2471.7
Terminal thiolate [His N2 Cu(II)SCys SMet ]+ 1.54 1 1 45 10.3 n/ae 2473.6
Olefinic enedithiolate [Ni(III)(dmedt)2 ]− 0.539 4 1 59 13.9 2474.5 2474.1
Aromatic enedithiolate (␲) [Ni(III)(bdt)2 ]− 0.799 4 1 74 13.0 2474.6 2474.3
Conjugated enedithiolate(␴) [Cu(II)(mnt)2 ]2− 0.801 4 1 54 17.8 2474.0 2473.8
Conjugated enedithiolate(␲) [Ni(III)(mnt)2 ]− 1.05 4 1 60 21.1 2474.4 2473.8
Dithiocarbamate [Cu(II)(dtc)2 ] 1.09 4 1 48 27.4 2474.4 2474.3
a
Normalized analytical peak intensity.
b
Number of absorbers.
c
Number of electron holes.
d
Approximate energies of the edge jump inflection points for complexes (C) and ligands (L).
e
In stellacyanin metalloprotein, edge position from the coordinated thiolate is not resolved due to overlapping Cys and Met residues.

corresponds to 42% S 3p character from XPS measurements as ana-


lyzed by VBCI model [30]. The corresponding 4p/edge inflection
points are at 2473.6 eV for the complex and 2471.7 eV for the free
sulfide ligand. The S K-edge spectrum of the Na2 S salt has two
intense features in the 2471–2475 eV region (bottom of Fig. 3A);
however, the second intense feature can be explained by construc-
tive interference of the scattered and backscattered ionized S 1s
core electron (broad undulating EXAFS feature). In going from sul-
fide to the thiolate (bottom three spectra in Fig. 3) the edge position
is expected to shift toward higher energy due to the increase of the
Zeff (S) as the sulfur charge decreases. The sensitivity of the S K-edge
technique to ligand-based chemical shifts is well demonstrated by
the difference between an aliphatic and aromatic thiolates from the
spectra of the NaSPh and NaSEt salts.
Continuing to use independent experimental data from XAS
measurements, we can define additional entries in Table 2.
Analysis of 33 S and complementary metal hyperfine cou-
pling constants from EPR/ENDOR/ESEEM measurements for the
[Ni(III)(bdt)2 ]− [104], [Ni(III)(dmedt)2 ]− [17], [Cu(II)(mnt)2 ]2−
[7,105–109], [Ni(III)(mnt)2 ]− [7,17,18], and [Cu(II)(dtc)2 ] [110,111]
EPR and S K-edge XAS data defines additional transition dipole
integrals as shown in Table 2. In addition to the above experimen-
tally determined transition dipole integrals, we took into account
two additional data points that can be considered as examples for
the most oxidized S-ligands that can coordinate to a metal center
[112]. These provide critical constraints in developing a ligand-
based transition dipole relationship as they mark the upper limit
for ligand-based XAS shifts. In Fig. 4A, the coordination compound-
based S 1s → 3p dipole integrals (IC ) are plotted as a function of the
complexes’ edge inflection points (E0C ) relative to that of the Na2 S
ligand salt (Fig. 3A, 2471.7 eV).
Comparing the relative values of the S transition dipole integrals
for reference complexes with Cu(II), Ni(III), and Fe(III) metals from
Table 2, one can see a qualitative trend of increase in I values that
correspond with less negative S and thus higher Zeff (S). Previously
a linear dependence of the dipole integral I on Zeff was employed
as indicated by Mulliken population analysis [26,31] and by a more
sophisticated evaluation of oscillator strengths [33,53]. As an exam-
ple, the hollow diamonds in Fig. 4A mark the positions for the two
most oxidized sulfur ligands, ROS− and RO2 S− , and indicate that
their corresponding IC values calculated from a linear relationship.
However, the smaller diamond symbols indicate considerable devi-
ations from a linear trend, especially when the mnt2− and the dtc−
complexes are considered. This non-linearity is unexpected on the
basis of past literature of S K-edge XAS; however, it can be rational-
ized by considering the cumulative effects of the ligand-based and
coordination-based chemical shifts in spectroscopic features due to
change in Zeff (S) as a function of Ni S coordination in addition to
differences in S oxidation states and C S bonding. The ligand-based Fig. 4. Correlation of edge positions relative to Na2 S edge-inflection point
chemical shifts for the uncoordinated S-ligands are illustrated by (2471.7 eV) and S 1s → 3p dipole integrals (A) E0C versus IC (B) graphical illustration
the spectroscopic changes in Fig. 3. of the development of ligand-based dipole integrals (C) E0L versus IL .
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 573

To further explore the relationship between I and Zeff (S) as As mentioned above, the data for the ROS− and RO2 S− ligands
expressed by experimental edge positions, we need to separate [113,114] provide a critical upper boundary condition for establish-
the effect of coordination-based chemical shift. As the free ionic ing a relationship between Zeff (S) and the dipole integral. Another
S-ligand coordinates to the metal complex, L → M electron dona- reference point as a lower boundary condition is the data for the
tion takes place (hence the covalent M S bonds) and the S-ligand Na2 S, which represents the most reduced form of a S-ligand. For the
becomes partially oxidized. The most affected part of the ligand will specific mathematical relationship for IL (E0L ), we propose the use of
be the sulfur due to its direct bonding to a metal center and thus the a second order polynomial for considering the Zeff dependence of
Zeff (S) will increase. Fig. 4B represents the illustration for connect- the S 1s → 3p transition dipole integral, since we are using energy
ing the coordination compound-based dipole integrals (IC ) with positions (E0L or E0C ) as the measure of effective nuclear charge.
the relative edge positions of the uncoordinated S-ligands (E0L ). The According to the Bohr model [115] with the extension by Mose-
red vertical error bars represents the range for each ligand-based ley [116] that considers electron shielding and use the concept of
dipole integral (IL ) with the maximum values being IC for a given effective nuclear charge, both the donor core and acceptor valence
ligand/complex. The blue diamonds mark the IC values and corre- electron binding energies are proportional to Zeff 2 . This is actually
sponding edge inflection points for the coordination compounds not a novel idea, since similar relationships have already been dis-
(E0C ). The purple arrows indicate the possible directions for going cussed in literature for a series of first row transition metal K-edges
from the experimental, coordination compound-based (E0C , IC ) data [117–120]. Distortion from linearity can already be observed in the
points to the corresponding ligand-based (E0L , IL ) points. The slope free ligand spectra of Fig. 3, since as the Zeff (S) increases, the highest
of these arrows will be used later to define the change in the transi- energy, resolved bound state excitations overlapping with the ion-
tion dipole integral (IC –IL ) for a given coordination-based chemical ization threshold. Using the relative energy scale (E0L ) and taking
shift (E0C –E0L ). For the most oxidized MeSOR and MeSO2 R ligands, the into account the reference Na2 S being the most reduced S-ligand,
arrows also point toward greater IL indicating that we will allow we constrained the second order relationship as shown in Eq. (5)
for deviation with positive and negative slopes. The hollow blue 2
rectangles show that acceptable range for the dipole integrals as I L = a(E0L ) + b (E0L ) + c (5)
defined by electronic structure calculations [112–114].
to b = 0, which minimizes the ligand-based dipole integral at the sul-
The experimental data for the [Cu(II)(mnt)2 ]2− and
fide. The maximum value for ‘c’ is the IC for the [FeS2 ]− (8.90 eV). In
[Ni(III)(mnt)2 ]− complexes from Table 2 provide an opportu-
this case the slope parameter for sulfide would be zero (horizontal
nity to analytically determine a slope parameter, as introduced
arrow in Fig. 4B). The ‘a’ parameter in Eq. (5) becomes constrained
above, for taking into account the XAS chemical shift due to Ni S
by the theoretical estimate [112] of the orbital composition of
bond formation. Using the energy difference between the edge
[Ru(II)-SOR]+ and [Ru(II)-SO2 R]+ complexes (hollow blue box sym-
positions for the ionic free ligand salt and the complexes, we
bols in Fig. 4B). In order to extend the known slope parameter for
can estimate a free ionic ligand-based dipole integral (IL ) for a
mnt2− ligand defined in Eq. (3) to the entire series, we evaluated
hypothetical free mnt2− ligand that is not involved in any covalent
both second order and linear relationships and the data for the oxi-
interaction. Eqs. (3a) and (3b) utilize a relationship between the
dized S-ligands ruled out the former. By simultaneously varying the
coordination-based chemical shift (E0C –E0L ) and the experimental
‘a’ and ‘c’ parameters from Eq. (5), linking the intercept of the linear
IC , such as IC = IL + slope × (E0C –E0L ). relationship between the relative edge inflection point and slope
parameter to the slope value determined for the Na2 S ligand (low-
[Cu(II)(mnt)2 ]2− : 17.8 eV = I L (mnt2− )
est energy set of arrows in Fig. 4B) we developed a favorable set of
+ ␴-slope · (2474.0 − 2473.8 eV) (3a) parameters that satisfy all boundary conditions considered above.
The resulting uncoordinated S-ligand-based transition dipole inte-
[Ni(III)(mnt)2 ]− : 21.1 eV = I L (mnt2− ) grals as a function of relative edge inflection point are plotted in
Fig. 4C.
+ ␴/␲-slope · (2474.4 − 2473.8 eV) (3b) There are two data points marked with teal dots in Fig. 4C
that are not on the green dashed lines. These correspond with lig-
Despite the expected difference between the ␴ and ␲ M L over- ands that contain N atoms, which are conjugated in both ␴- and
lap, from the relative pre-edge intensities in [Ni(III)(mnt)2 ]− we ␲-orbitals with the S donor orbitals. In both cases, greater flu-
can approximate that the Ni(III) receives about the same amount orescence intensity is observed in XAS measurements than the
of electron density (∼0.5 e− /hole) from the S ligands for both ␲- and corresponding unsubstituted hydrocarbon ligands and thus the
␴-holes. The 0.2 eV shift for the Cu(II) complex with a single elec- IL (E0L ) relationship needed to be shifted up by about 10 eV. This
tron hole in the d-manifold can be correlated with the 0.6 eV shift enhanced fluorescence signal is also expected for pre-edge features
for the Ni(III) complex with three electron holes. The difference of for Ni(II)/Cu(II) thiourea and thiohistidine complexes. In the future,
Eqs. (3b) and (3a) becomes we plan to evaluate the cause of the excess fluorescence emission
channels relative to corresponding unsubstituted ligands by reso-
16.5 = ␴-slope + ␲-slope (4) nant inelastic X-ray scattering at the ligand K-edge energy region
From a comprehensive TDDFT study for a large set of complexes [121].
of the bdt2− ligand [53], a ␴/␲ slope ratio was obtained to be 1.2.
This would give a value of 16.0 eV for IL (mnt2− ) (Fig. 4B, teal dot). 3.4. Experimental S 3p contributions to the Ni S bonds in
Variation of the ␴/␲ slope ratio in the range of 0.5–2.0 would not [Ni(II)S4 ] complexes
affect the IL (mnt2− ) by more than 0.7 eV (4%). Thus, we can con-
sider the 16.0 ± 0.7 eV to be a reasonable estimate for the free ionic The estimated free ligand-based transition dipole integrals from
ligand-based S 1s → 3p transition dipole integral for the mnt2− Fig. 4C can now be used to extrapolate to other ligands with
ligand. Due to the lack of the complementary XAS data for any other unknown dipole integrals. The second order IL (E0L ) relationship for
ligands considered here, this is the only point that marks the ligand- the uncoordinated S-ligand provides an estimate for the transition
based Zeff (S) as expressed by E0L and the corresponding free ionic dipole integral for the ligand without being involved in any covalent
ligand-based dipole integral for S 1s → 3p core excitation (IL ) that interaction. The linear slope (E0C –E0L ) relationship defines the change
is known to date. needed for extrapolating to IC from IL . Table 3 summarizes the list
574 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

Table 3
Summary of parameters used to estimate coordination compound-based dipole integrals (IC ) for the studied [Ni(II)S4 ] complexes.

Complexes E0L , eVa IL , eV E, eVb Slopec IC , eV S 3p, %


 2− d
[Ni(SPh )4 ] 2473.6 5.8 0.4 13.9 11.4 30
[Ni(nbdt)2 ]2− 2473.1 5.2 1.0 8.6 13.8 56
[Ni(dmedt)2 ]2− 2474.1 6.6e 0.7 20.7 21.2 52
[Ni(bdt)2 ]2− 2474.3 7.0 0.5 23.9 19.0 27
[Ni(CNedt)2 ]2− 2474.0f 6.4 0.7 18.5 18.4 45
[Ni(mnt)2 ]2− 2473.8 16.0 0.5 16.4 24.2 35
[Ni(dtc)2 ] 2474.3 16.9 0.6 23.9 31.2 31
[Ni(ttctd)]2+ 2475.5 9.8 0.3 48.5 24.4 39
a
Approximate S 1s ionization threshold from Fig. 1B for the complex and Fig. 3B for the ligand.
b
Chemical shift from the edge inflection point between the coordinated and the free ionic S-ligand as a result of M S bonding.
c
The slope values for Eq. (4) are estimated from the minimal correction to reach the upper limit of dipole integral for a given or similar S-ligand (Fig. 4C).
d
The chemical shift was determined from the free ligand salt shift in Figure S2.
e
The Na2 edt spectrum was used.
f
The average edge inflection point of Na2 edt and Na2 mnt was used.

of parameters used to obtain the transition dipole integrals for the interesting exception for the [Ni(dtc)2 ] complex, where increase in
Ni complexes (IC ) from their corresponding free ligand value (IL ). the HF mixing (such as BHandHLYP with 50% HF exchange) gives
The error bar or uncertainty range of about 5–7% can be assumed the upper limit of 68% S character instead of an expected lower
for the dipole integral values due to uncertainties in data collec- limit. This implicates an inverted bonding description in the the-
tion, normalization, and fitting. There are two exceptions to this as oretical electronic structure, where the electron from the ligand
noted by footnotes ‘d’ to ‘f’ in Table 3. The comparison of free and has already been transferred to the Ni(II) to reduce it to a formal
coordinated aryl thiolate ligands posed a significant challenge as Ni(0) state and the two radicalized, effectively neutral dtc− ligands
documented in Figure S2 and instead of using the edge-inflection become antiferromagnetically coupled through the Ni center. As
points for the coordination-based chemical shift (1.5 eV) we con- observed before [13], the magnitude of the coupling versus cova-
sidered the energy shift used to subtract the free-ligand spectrum lent bond is determined by the particular hybrid functional used.
from the complex’s spectrum (0.4 eV). Furthermore, we did not The metaGGA and hybrid metaGGA functionals give similar aver-
have data for the corresponding free S-ligand of [Ni(dmedt)2 ]2− , age results to the GGA and hybrid functionals, respectively, but with
[Ni(CNedt)2 ]2− ; thus some straightforward approximations had to somewhat larger variation.
be made (see Table 3 footnotes). The comparison of the calculated overall S contribution from
Table 4 and the experimentally determined S 3p characters from
3.5. Theoretical electronic structure of [Ni(II)S4 ] complexes using Table 3 reveals a few mentionable differences. A key feature of
DFT the calculated values that the LUMOs in all square planar com-
plexes have about the same S character (43–51% in AIM) and this
A benefit of deriving ground state bonding descriptions from is considerably more than in the tetrahedral tetrathiolate com-
XAS measurements is the direct evaluation of theoretical orbital plex (29% in AIM). Judging from the large variation of the pre-edge
compositions from the popular density functional theory-based intensities (Fig. 2) and the reasonable range for estimated tran-
electronic structure calculations. Table 4 summarizes the func- sition dipole integrals (Fig. 4C) this similarity among the square
tional dependence of the sulfur character in the LUMOs using a planar [Ni(II)S4 ] complexes is not reproduced by the S K-edge data.
theoretically converged basis set and three population analysis Considering a reverse approach that is described in Section 3.3 for
methods. Despite some uncertainties in the ground state spin state Fig. 4, we can take the calculated overall S contributions to the
for Cu(II) complexes as a function of density functionals [40,41], all LUMO and estimate a DFT-based transition dipole integral (IDFT in
calculation in Table 4 reproduced the ground spin state for the tetra- Table 4). However, the IDFT values mostly underestimate the rea-
hedral and square planar complexes to be S = 1 and 0, respectively sonable range of dipole integrals from Fig. 4C and Table 3, which
(Table S2). is consistent with the general expectation that pure or gradient
It is also important to note that only the Mulliken (MPA) and corrected DFT functionals generally give too covalent ground state
Weinhold’s natural (NPA) population analysis can give S 3p orbital orbital picture relative to experiment [34,39].
compositions that are directly comparable to experiment, while
Bader’s atoms in molecules (AIM) method provides total sulfur 4. Discussion
atomic contributions. The role of the S 3s is negligible for these
complexes, thus the results of the latter can be directly compared With the given systematic study of [Ni(II)S4 ] homoleptic coor-
with the former methods. The range of calculated S contributions dination complexes we extended the series of classical S-ligands,
to the LUMO varies remarkably as a function of population analy- such as thiolates, aliphatic dithiolates, and enedithiolates to some
sis method. Generally the MPA method gives the largest deviation of the less reduced forms including conjugated dithiocarbamates
from the more comparable NPA and AIM results. In the discussion and thioethers. The XAS data collection and data reduction at the
below, we wish to mainly focus on the AIM results due to our past sulfur K-edge are straightforward for these complexes. The pre-
success in using this more advanced population analysis methods edge and rising-edge features can be readily assigned with little
in describing electronic structure of Fe–S clusters [98]. uncertainty especially given the possibility of comparing spectro-
The GGA functionals provide generally a narrow distribution of scopic features for a broad series of uncoordinated ligands and their
the LUMO’s S character of the Ni(II)S4 complexes. The hybrid GGA Ni(II) complexes. All square planar complexes can be characterized
functionals due to the variation of the amount of HF mixing gave a by intense pre-edge features except for the [Ni(bdt)2 ]2− complex
considerably broader range with S contributions that are below the (0.9 eV), which is in between the pre-edge areas observed for the
GGA values. This is consistent with the expected effect of mixing HF tetrahedral (0.5 eV) and square planar complexes (considerably
exchange with exchange density functional and thus emphasizing greater than 1 eV). However, the conversion of the pre-edge inten-
ionic versus covalent character in Ni S bonding. We found here an sities to actual orbital composition is not straightforward. Despite
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 575

Density functional and population analysis method dependence of the Ni S bonding as expressed by total S contribution (%) to the LUMO for the studied [Ni(II)S4 ] complexes (MPA = Mulliken, NPA = Weinhold’s natural orbital,

Combined uncertainty due to the variation of sulfur orbital composition as a function of density functionals, population analysis method (total for AIM and S 3p for MPA and NPA) and uncertainty in H atomic positions of the
attempts to find a useful linear relationship between either the

IDFT , eV
complex (E0C ) or the ligand (E0L )-based transition dipole integrals

AIM

11
17
23
10
26
13
21
we had to consider a second order relationship between the Zeff (S)
and the dipole integrals. We carried out this by using an experimen-

16
tal energy position (white-line position, rising-edge or edge-jump

9
9
8
5
6

5
±
±
±
±
±
±
±
inflection point) for the measure of Zeff , which is the function of
AIM

28
46
49
50
43
51
46
2 as the experimental measure of sulfur effective nuclear charge.
Zeff
The non-linearity is further supported by the experimental obser-
11

17
14
12
vation (Fig. 3) that as the sulfur effective nuclear charge increases;
9

5
5
±
±
±
±
±
±
±
NPA

the core level excitations go up in energy to such an extent that


28
48
55
64
36
53
53
they overlap with the ionization threshold of the core electron. We
found that the first inflection point (E0 ) of the XAS spectrum after
the last resolved (S 1s → C S ␴*) feature gives a reasonable exper-
Overalle

11
5
4
6

9
9
3
MPA

±
±
±
±
±
±
±
imental estimate for the Zeff (S) in both the complexes and in the
22
51
53
60
49
54
49

free ligands. This is well represented by the shifts of spectroscopic


features of the S-ligand series in Fig. 3 without the presence of sig-
nificant metal–sulfur covalent bonding (except for the Li salts or
16–26
39–54

59–61
40–48

40–45
AIM

protonated ligands). For the sulfide ligand the first intense feature
48
37

along the rising edge will need to be considered as the measure


of the Zeff (S) due to the lack of S C bonding. Using the currently
23–25

43–63
12–39
11–57
11–55

available data from the literature, we can set up a parabolic rela-


Hybrid metaGGAd

9–51
NPA

tionship for the free ionic ligand-based S 1s → 3p transition dipole


52

integrals as a function of edge position of the S-ligand salt (E0L ). This


relationship was obtained from converting the EPR and XPS-based
21–22
51–39
42–57
56–91

45–52
40–48

transition dipole integrals (IC at E0C , Table 2) for various M S com-


MPA

38

plexes to corresponding free ionic S-ligand-based integrals (IL at E0L ,


Fig. 4C) by estimating the coordination-based chemical shift in the
B1LYP [71,82], BHandHLYP [71,84], mPW1PBE [73,76], MPW1LYP [87], B3LYP [71,83], O3LYP [71,88], MPW1K [85], MPW3LYP [86].

Zeff (S) due to the formation of the M S bonds. Considering various


47–49

48–65
45–47
30–48

48–50
AIM

boundary conditions obtained from the S K-edge spectra of sulfide,


44
50

sulfinate, and sulfonate ligands and their complexes as examples


for the most reduced and most oxidized S-ligands, we developed a
52–55
55–57

48–51
57–59
55–58

linear relationship to convert the coordination-based chemical shift


NPA

33

57

to a correction of ligand-based integrals to obtain complex-based


integrals.
metaGGAc

As described earlier [3], the tetrathiolatonickel(II) complex can


24–25
50–53
55–57

46–48
52–53
47–51
MPA

be regarded as classical coordination compound in which the par-


56

tially occupied orbitals and thus the M L bond are dominantly


metal-based. Therefore, any redox chemistry will primarily affect
the metal. The previous interpretation of XAS data using the uncor-
40–49
40–52
46–52
38–46
46–68
44–49
22–30
AIM

rected IC (Cu-SEt) dipole integral gave 33% S character for each of


the LUMOs, which we estimate to be somewhat less (30%). How-
ever, in going from terminal thiolates in tetrahedral coordination
4–35
36–56
39–59
57–49
14–52
52–61
42–60

environment to covalently linked thiolates in the square planar


BLYP [35,71], BP86 [35,72], mPW [73], OLYP [71,74,75], PBE [76], PW91 [77,78].
NPA

[Ni(nbdt)2 ]2− complex the pre-edge intensities go up from 0.56


Hybrid GGAc

to 1.28 eV (Table 1). Even with the increased dipole moment (IC
from Table 3) from 12.2 to 15.0 eV, the S 3p contribution goes up
43–54
42–58
51–57
41–49
48–54
46–52
10–26

in the square planar complex to be more than 50%. This is an indi-


MPA

cation of inverted bonding, where the LUMOs become S-based and


as a result, the electrophilic site shifts from the Ni to the S atoms.
In a valence bond/ionic limit description, the formally Ni(II) cen-
33–32
49–51

51–52

50–51

ter is effectively reduced to Ni(0) and the formally RS2 2− ligands


M06 [89], M062X [89], M06-HF [90,91], TPSSh [80].
AIM

52

46

49

become radicalized anions (RS2 −• ). This is a similar description to


the dianionic, olefinic enedithiolate complex [13]. Our experimen-
AIM = Bader’s ‘atoms in molecules’ methods).

tal S 3p values for the [Ni(nbdt)2 ]2− and [Ni(dmedt)2 ]2− complexes
36–37
56–54
58–59
58–59
52–53

M06L [79], TPSS [80], TPSSLYP1W [81].


NPA

are practically identical. This suggests that by covalently linking


61
60

two donor atoms in 1,2 position either by a strained ligand sys-


tem (as in nbdt2− ) or with a double bond (as in dmedt2− ) the
52–54
58–59
57–59
49–72
55–87
52–53

S ↔ S repulsion within and in between the ligands can destabilize


GGAb

MPA

26

in energy the S-based donor orbitals above the acceptor metal d-


crystal structures used.

orbital and thus trigger intramolecular electron-transfer. The latter


is well facilitated by the efficient Ni S overlap in the square planar
[Ni(dmedt)2 ]2−
[Ni(nbdt)2 ]2−

complex relative to that in a tetrahedral complex. Using the satu-


[Ni(SPh )4 ]2−

[Ni(mnt)2 ]2−

[Ni(ttctd)]2+
[Ni(bdt)2 ]2−

rated def2-TZVP basis set with respect to LUMO composition, the


[Ni(dtc)2 ]

overlap integral as calculated from the product of overlap and den-


Table 4

sity matrices drops from 0.086 to 0.006 per Ni S bond in going from
square planar to tetrahedral complexes. This is dominantly due to
b

e
a
576 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

the 0.1 Å elongation of the Ni S bonds in the latter in addition to Equally interesting is the case of the Ni(II)-thioether complex.
the non-ideal alignments of the S donor and vacant Ni 3d orbitals. Compositionally it is a unique Ni(II) coordination compound, since
Considering the [Ni(dmedt)2 ]2− complex as a starting point for normally thioether does not coordinate to late transition metals.
the series of enedithiolate complexes shown in Table 3, we can This is due to the limited nucleophilicity of the S centers despite
describe a chemically meaningful trend about Ni S ␴-bonding on the availability of S lone pair for coordination, which also means
the basis of experimental S covalency from XAS. The LUMO in the that the ionic interactions between the Ni(II) and the S centers
[Ni(dmedt)2 ]2− complex has about 52% S 3p character, which corre- were diminished relative to the anionic ligands. However, the
sponds to inverted bonding. Substituting one of the methyl groups ideal geometry of the tetrathiocrownether (ttctd) allowing for 90◦
in dmedt2− with a nitrile group as in CNedt2− the S contributions S Ni S bite angles and the vicinity of the four sulfur donors allow
drops to about 45%. The second substitution leading to the mnt2− for the formation of the Ni(II) complex. The Ni S overlap is effi-
ligand further drops the S character to 35%. This trend continues cient as also witnessed by the strong color of the [Ni(ttctd)]2+
for the [Ni(bdt)2 ]2− complex. This series nicely demonstrates the complex due to the presence of a S → Ni charge transfer band
importance of correcting the transition dipole integral due to the with molar extinction coefficient of 370 dm3 mol−1 cm−1 in GBL
large variation in Zeff (S) between various S-ligands and furthermore [22]. As mentioned in the introduction, this complex displays
the smaller adjustment as a result of metal coordination. ligand-exchange-based thermochromic behavior in the 25–125 ◦ C
Keeping this in mind, it is then not surprising that for the range. The estimated Hrxn of 60 kJ mol−1 for the ligand-exchange
[Ni(bdt)2 ]2− complex we obtain such low S 3p values of about 27%. reaction (Eq. (1)) suggests that the Ni S bond is only moder-
The EPR data used for the transition dipole integral and the free ately covalent and has limited ionic character. Otherwise the
bdt2− ligand salt define a exceptionally high S contribution (74%) reversible ligand-exchange thermochromism would not take place
for the out-of-plane ␲-orbital [104]. The corresponding S K-edge for [Ni(II)(ttctd)]2+ as can be observed for all the enedithiolates, for
feature [53] has an intensity of 0.801 eV, while the adjacent fea- example. This means that the experimental value of about 39% is a
ture from the in-plane ␴-orbital with two electron holes has only reasonable number for the S 3p character of the Ni S bonds. This is
1.09 eV. Thus, the latter corresponds to about 0.545 eV intensity per actually a considerable electron donation from a neutral S-ligand
electron hole, which is about 40% overall reduction of S covalency to the Ni(II) in comparison to the similar S 3p values for the anionic
per hole. Even without a correction to the dipole integral for ␴- tetrahedral tetrathiolate complex.
versus ␴ + ␲-bonding the expected S character for the Ni(III) com- Without going into specific details for the performance of
plex in the ␴-hole is about 44%. The missing 17% S character from each and every density functional, we found that the previously
our estimated value for the Ni(II) complex is likely due to the use of described [34,38,39] transferability of a specific hybrid functional
the already covalent Li+ salt of the bdt2− free ligand. The protonated does not hold for [Ni(II)S4 ] complexes. The most peculiar finding
H2 bdt XAS data gives an even larger deviation from the expected is that all square planar [Ni(II)S4 ] complexes were described by
S character due to the higher energy position of the edge jump for a greatly covalent Ni S bonding with close to 50% S character.
the free ligand spectrum (Figure S12). While these values are reasonable for the enedithiolates and even
The dithiocarbamate and the thioether ligands, as the two least the constrained 1,2-dithiolate ligands, it certainly deviates consid-
reduced S-ligands of the studied series, showed the most surpris- erably even from the upper limit of XAS values for the dtc− and
ing S 3p characters for the experimentally probed LUMOs. The dtc− the ttctd complexes. As mentioned above, we need to be cautious
ligand is particularly interesting with respect of the entire S-ligand with directly comparing the experimental and the computational
series, since the intraligand S ↔ S repulsion is reduced by removing results, since all computational models were only embedded in
one electron relative to the dithiolate ligands despite the reduced an isotropic polarizable continuum model and none included the
S· · ·S distance. Furthermore, from electronic structure calculations potentially critical fine tuning effect of counter ions and adjacent
using the saturated def2-TZVP basis set, we found that in going molecules from the crystal packing environment. The systematic
from a 2.18 Å Ni S bond length to 2.20 Å, as in [Ni(dtme)2 ]2− rel- evaluation of these effects and the limitations of linear response
ative to [Ni(dtc)2 ], the Ni S orbital overlap decreases by 3% per theory-based TDDFT calculations are being investigated.
sulfur atom; while moving the sulfur atoms off axis by about 10◦
also decreased the overlap by about 4% per sulfur atom. As a conse-
quence, while the [Cu(dtc)2 ] complex in Table 2 was characterized 5. Conclusions
by covalent bonding of about 48% S in the paramagnetic SUMO
from EPR, the estimates for the [Ni(dtc)2 ] complex from XAS is Despite some uncertainties of employing quantitative XANES
31% (Table 3) despite the more intense pre-edge features (1.09 analysis for the S K-edge spectra of a spectrochemical series for
versus 1.60 eV, respectively). This is due to the non-linearity of the [Ni(II)S4 ] complexes, we can still consider the experimental deter-
transition dipole integral as it considerably increases relative to mination of the composition of unoccupied frontier orbitals from
the thiolates and enedithiolates. From Table 2, the IC values for pre-edge and rising-edge features to be a very powerful approach
Cu(II) complex with thiolate, enedithiolate, and dithiocarbamate that is worthwhile to continue developing. The current limitations
changes from 10.3, 17.8, to 27.4 eV, respectively. Due to the open- of this approach are mainly due to the limited complementary data
ing of the additional electron hole in the Ni(II) complex relative to sets for comparison and critical assessment of the data dependence
the Cu(II), the energy of the edge position shifts up by 0.5 eV or of detection method and detector instrument, sample preparation
0.6 eV relative to the Cu(II) complex and the free ligand, respec- and thickness, and data normalization procedures. Furthermore,
tively. These chemical shifts with a ␴-slope value of about 24 from the experimentalists need to be critical in evaluating the peak
Table 3 give an IC value that is above 30 eV, which defines about fitting procedure since it can have a dramatic influence on the
31% S 3p character in the LUMO. The difference between the above resulting covalency parameters. The computed electronic struc-
two complexes suggests that we need to expect considerable bond- tures from the popular density functional theory showed serious
ing differences between the dithiocarbamate-coordinated Cu(II) discrepancies among each other as well as to the experimental XAS
and Ni(II) complexes, where Cu(II) has an inverted bonding, while data. Therefore, one should not rely too heavily on any one func-
Ni(II) displays classical coordination chemistry. By extending the tional or population analysis model. It is beyond any doubt that
series with additional metal complexes, such as Zn(II), Fe(II)/Fe(III), complementary experimental techniques and internal checks are
Mo(IV), we are currently investigating a broad family of dtc needed for XAS data to obtain a reasonable experimental descrip-
complexes [111]. tion of metal–ligand bond covalency.
M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578 577

The already available XAS and EPR data allow for a yet to be absorption spectra with the raw, normalized, and fit data can be
explored phenomena between the composition of the ligands and downloaded at http://computational.chemistry.montana.edu/SI.
the magnitude of transition dipole integral. The notably higher Supplementary data associated with this article can be found, in
transition dipole integrals in Table 3 and Fig. 4C (teal dots) cor- the online version, at http://dx.doi.org/10.1016/j.ccr.2012.07.020.
respond with ligands that contain N substituents electronically
interacting with the sulfur centers. The intensity enhancement of References
spectroscopic features by a heteroatom is the subject of our future
XAS and RIXS studies. [1] A.B.P. Lever, Inorganic Electronic Spectroscopy, Elsevier, Amsterdam, 1984.
Using the available experimental data we can classify the [2] A. Silver, S.A. Koch, M. Millar, Inorg. Chim. Acta 205 (1993) 9.
[3] K.R. Williams, B. Hedman, K.O. Hodgson, E.I. Solomon, Inorg. Chim. Acta 263
series of [Ni(II)S4 ] homoleptic complexes into three categories.
(1997) 315.
The tetrahedral tetrathiolate complex can be described with clas- [4] J.A. McCleverty, Metal 1, 2-Dithiolene and Related Complexes, John Wiley &
sical coordination chemical bonding description with the lowest Sons, Inc., Hoboken, NJ, USA, 1968.
[5] G.N. Schrauzer, Acc. Chem. Res. 2 (1969) 72.
unoccupied frontier orbitals being metal based. Constraining two
[6] R. Eisenberg, Structural Systematics of 1,1- and 1, 2-Dithiolato Chelates, John
negatively charged thiolates either by a ring system (nbdt2− ) or Wiley & Sons, Inc., Hoboken, NJ, USA, 1970.
double bonds (dmedt2− ) induces inverted bonding description, [7] A.H. Maki, A. Davison, N. Edelstein, R.H. Holm, J. Am. Chem. Soc. 86 (1964)
where each ligand donates more than 0.5 electron per electron hole 4580.
[8] S.I. Shupack, R. Williams, H.B. Gray, E. Billig, R.J.H. Clark, J. Am. Chem. Soc. 86
to the metal and thus effectively reduces the metal. In these dian- (1964) 4594.
ionic complexes, the formal divalent metal is better described by [9] G.N. Schrauzer, V.P. Mayweg, J. Am. Chem. Soc. 87 (1965) 1483.
a zero-valent metal that is stabilized by covalent and/or L M L [10] J.R. Baker, A. Hermann, R.M. Wing, J. Am. Chem. Soc. 93 (1971) 6486.
[11] K. Wang, E.I. Stiefel, Science 291 (2001) 106.
superexchange interactions. The substitution of the enedithiolate [12] V. Bachler, G. Olbrich, F. Neese, K. Wieghardt, Inorg. Chem. 41 (2002) 4179.
ligand framework with strongly electron withdrawing groups, such [13] R.K. Szilagyi, B.S. Lim, T. Glaser, R.H. Holm, B. Hedman, K.O. Hodgson, E.I.
as in CNedt2− , mnt2− , or bdt2− , attenuates the S → Ni electron dona- Solomon, J. Am. Chem. Soc. 125 (2003) 9158.
[14] W. Kaim, B. Schwederski, Coord. Chem. Rev. 254 (2010) 1580.
tion. The most unexpected finding is related to the two less reduced [15] R. Eisenberg, Coord. Chem. Rev. 255 (2011) 825.
S-ligands, since despite their intense pre-edge features, they are [16] R.K. Szilagyi, in: B.R. King (Ed.), Encyclopedia of Inorganic Chemistry, Second
best described by a covalent, but still normal bonding scenario. This edition (EIR-2), John Wiley & Sons, Ltd, 2009, p. 1.
[17] R.D. Schmitt, A.H. Maki, J. Am. Chem. Soc. 90 (1968) 2288.
is chemically quite reasonable, since in the [Ni(II)dtc2 ] complex the
[18] J.E. Huyett, S.B. Choudhury, D.M. Eichhorn, P.A. Bryngelson, M.J. Maroney, B.M.
vicinity of the two S donors, the difference in donor ability of the Hoffman, Inorg. Chem. 37 (1998) 1361.
1,1-dithiocarbamate versus a 1,2-dithiolate, and the reduced charge [19] B.S. Lim, D.V. Fomitchev, R.H. Holm, Inorg. Chem. 40 (2001) 4257.
[20] D.V. Fomitchev, B.S. Lim, R.H. Holm, Inorg. Chem. 40 (2001) 645.
per S donor cannot allow for the same efficient ligand-to-metal
[21] W. Rosen, D.H. Busch, J. Am. Chem. Soc. 91 (1969) 4694.
electron donation as in the enedithiolate complexes. Furthermore, [22] H.J. Byker, P.H. Ogburn, D.A. Vander Griend, B.S. Veldkamp, D.D. Winkle, US
the formation of the Ni(II) complex with the cyclic tetrathioether 7,542,196, Pleotint, L.L.C. (West Olive, MI) USA, 2009.
ligand is mainly possible due to the ideal S Ni overlap despite the [23] B.S. Veldkamp, D.A. Vander Griend, H.J. Byker, personal communication, West
Olive, MI, 2012.
considerably reduced nucleophilicity of the S centers and the avail- [24] E.I. Solomon, B. Hedman, K.O. Hodgson, A. Dey, R.K. Szilagyi, Coord. Chem.
ability of a single S lone pair for Ni S bond formation with only a Rev. 249 (2005) 97.
modest ionic character. [25] B. Hedman, K.O. Hodgson, E.I. Solomon, J. Am. Chem. Soc. 112 (1990) 1643.
[26] S.E. Shadle, B. Hedman, K.O. Hodgson, E.I. Solomon, J. Am. Chem. Soc. 117
(1995) 2259.
[27] F. Neese, B. Hedman, K.O. Hodgson, E.I. Solomon, Inorg. Chem. 38 (1999) 4854.
Acknowledgements [28] R.B. Boysen, R.K. Szilagyi, Inorg. Chim. Acta 361 (2008) 1047.
[29] S.E. Shadle, J.E. Pennerhahn, H.J. Schugar, B. Hedman, K.O. Hodgson, E.I.
Portions of this research were carried out at the Stanford Solomon, J. Am. Chem. Soc. 115 (1993) 767.
[30] K. Rose, S.E. Shadle, T. Glaser, S. de Vries, A. Cherepanov, G.W. Canters, B.
Synchrotron Radiation Lightsource (SSRL), a Directorate of SLAC Hedman, K.O. Hodgson, E.I. Solomon, J. Am. Chem. Soc. 121 (1999) 2353.
National Accelerator Laboratory and an Office of Science User [31] Y. Izumi, T. Glaser, K. Rose, J. McMaster, P. Basu, J.H. Enemark, B. Hedman, K.O.
Facility operated for the U.S. Department of Energy Office of Sci- Hodgson, E.I. Solomon, J. Am. Chem. Soc. 121 (1999) 10035.
[32] S.D. George, L. Basumallick, R.K. Szilagyi, D.W. Randall, M.G. Hill, A.M. Nersis-
ence by Stanford University. The SSRL Structural Molecular Biology
sian, J.S. Valentine, B. Hedman, K.O. Hodgson, E.I. Solomon, J. Am. Chem. Soc.
Program is supported by the DOE Office of Biological and Environ- 125 (2003) 11314.
mental Research, and by the National Institutes of Health, National [33] R. Sarangi, S.D. George, D.J. Rudd, R.K. Szilagyi, X. Ribas, C. Rovira, M. Almeida,
K.O. Hodgson, B. Hedman, E.I. Solomon, J. Am. Chem. Soc. 129 (2007) 2316.
Center for Research Resources, Biomedical Technology Program
[34] R.K. Szilagyi, M. Metz, E.I. Solomon, J. Phys. Chem. A 106 (2002) 2994.
(P41RR001209). Funding from the National Science Foundation to [35] A.D. Becke, Phys. Rev. A 38 (1988) 3098.
RKS (MCB 0744820 and CBET 0755676) and the NASA Astrobiology [36] A.A. Gewirth, S.L. Cohen, H.J. Schugar, E.I. Solomon, Inorg. Chem. 26 (1987)
Institute (NNA08C-N85A) are greatly appreciated. We also thank 1133.
[37] S.V. Didziulis, S.L. Cohen, A.A. Gewirth, E.I. Solomon, J. Am. Chem. Soc. 110
Prof. Serena DeBeer for providing us with the raw data for the (1988) 250.
[Ni(II)(bdt)2 ]2− complex. We would like to express our apprecia- [38] R.K. Szilagyi, E.I. Solomon, Curr. Opin. Chem. Biol. 6 (2002) 250.
tion to Prof. Pierre Kennepohl at University of British Columbia and [39] E.I. Solomon, R.K. Szilagyi, S.D. George, L. Basumallick, Chem. Rev. 104 (2004)
419.
Dr. Ritimukta Sarangi at SSRL for participating in an open review. [40] D. Rokhsana, D.M. Dooley, R.K. Szilagyi, J. Am. Chem. Soc. 128 (2006) 15550.
Their critical comments have greatly improved the manuscript and [41] D. Rokhsana, D.M. Dooley, R.K. Szilagyi, J. Biol. Inorg. Chem. 13 (2008) 371.
the applicability of the proposed method in estimating S 1s → 3p [42] D. Rokhsana, A.E. Howells, D.M. Dooley, R.K. Szilagyi, Inorg. Chem. 51 (2012)
3513.
transition dipole integrals. [43] E.I. Solomon, S.I. Gorelsky, A. Dey, J. Comput. Chem. 27 (2006) 1415.
[44] E.I. Solomon, Inorg. Chem. 45 (2006) 8012.
[45] R. Sarangi, S.I. Gorelsky, L. Basumallick, H.J. Hwang, R.C. Pratt, T.D.P. Stack, Y.
Appendix A. Supplementary data Lu, K.O. Hodgson, B. Hedman, E.I. Solomon, J. Am. Chem. Soc. 130 (2008) 3866.
[46] R.K. Szilagyi, M.A. Winslow, J. Comput. Chem. 27 (2006) 1385.
[47] T.V. Harris, R.K. Szilagyi, Inorg. Chem. 50 (2011) 4811.
Extended energy range of Figs. 1 and 3, details of rising-edge and
[48] T.V. Harris, R.K. Szilagyi, in: M. Ribbe (Ed.), Methods in Molecular Biology:
edge-jump subtraction, the atomic coordinates of computational Nitrogen Fixation, Springer, 2011, p. 267.
models in X-mol format are provided. A concise documenta- [49] A.S. Pandey, T.V. Harris, L.J. Giles, J.W. Peters, R.K. Szilagyi, J. Am. Chem. Soc.
tion about the derivation of the complex-based dipole integral 130 (2008) 4533.
[50] J.-H. So, P. Boudjouk, H.H. Hong, W.P. Weber, in: R.N. Grimes (Ed.), Inorganic
for Cu(dtc)2 is presented [111]. Larger electronic files, format- Synthesis, John Wiley & Sons, Inc., Hoboken, NJ, USA, 1992, p. 30.
ted checkpoint files, ADRP spreadsheets of all S K-edge X-ray [51] A.R. Hendrickson, R.L. Martin, N.M. Rohde, Inorg. Chem. 14 (1975) 2980.
578 M.S. Queen et al. / Coordination Chemistry Reviews 257 (2013) 564–578

[52] S. Fox, Y. Wang, A. Silver, M. Millar, J. Am. Chem. Soc. 112 (1990) 3218. [85] B.J. Lynch, P.L. Fast, M. Harris, D.G. Truhlar, J. Phys. Chem. A 104 (2000)
[53] K. Ray, S.D. George, E.I. Solomon, K. Wieghardt, F. Neese, Chem.-Eur. J. 13 4811.
(2007) 2783. [86] Y. Zhao, D.G. Truhlar, J. Phys. Chem. A 108 (2004) 6908.
[54] R.K. Szilagyi, P. Frank, S.D. George, B. Hedman, K.O. Hodgson, Inorg. Chem. 43 [87] N.E. Schultz, Y. Zhao, D.G. Truhlar, J. Phys. Chem. A 109 (2005) 11127.
(2004) 8318. [88] A.J. Cohen, N.C. Handy, Mol. Phys. 99 (2001) 607.
[55] I.J. Pickering, G.N. George, E.Y. Yu, D.C. Brune, C. Tuschak, J. Overmann, J.T. [89] Y. Zhao, D.G. Truhlar, Theor. Chem. Acc. 120 (2008) 215.
Beatty, R.C. Prince, Biochemistry 40 (2001) 8138. [90] Y. Zhao, D.G. Truhlar, J. Phys. Chem. A 110 (2006) 13126.
[56] D.J. Gardenghi, Automated Data Reduction Protocol, Version 1.1, [91] Y. Zhao, D.G. Truhlar, J. Phys. Chem. A 110 (2006) 5121.
http://computational.chemistry.montana.edu/ADRP/, Bozeman, MT, 2011. [92] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 7 (2005) 3297.
[57] M.J.T. Frisch, G.W., H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. [93] D. Feller, J. Comput. Chem. 17 (1996) 1571.
Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, [94] K.L. Schuchardt, B.T. Didier, T. Elsethagen, L. Sun, V. Gurumoorthi, J. Chase, J.
X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Li, T.L. Windus, J. Chem. Inf. Model. 47 (2007) 1045.
Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, [95] R.F.W. Bader, Acc. Chem. Res. 18 (1985) 9.
Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. [96] F. Corts-Guzman, R.F.W. Bader, Coord. Chem. Rev. 249 (2005) 633.
Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. [97] R.F.W. Bader, J. Phys. Chem. A 114 (2010) 7431.
Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. [98] L.J. Giles, A. Grigoropoulos, R.K. Szilagyi, Eur. J. Inorg. Chem. (2011) 2677.
Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, [99] T.A. Keith, AIMAll (Version 11.12.19), TK Gristmill Software, Overland Park,
C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. KS, USA, 2011.
Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, [100] R.S. Mulliken, J. Chem. Phys. 23 (1955) 1833.
G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B. [101] J.P. Foster, F. Weinhold, J. Am. Chem. Soc. 102 (1980) 7211.
Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Revision C.1, Gaussian, [102] S.E. Shadle, B. Hedman, K.O. Hodgson, E.I. Solomon, Inorg. Chem. 33 (1994)
Inc., Wallingford, CT, 2009. 4235.
[58] S. Miertus, E. Scrocco, J. Tomasi, Chem. Phys. 55 (1981) 117. [103] S.J. George, M.D. Lowery, E.I. Solomon, S.P. Cramer, J. Am. Chem. Soc. 115
[59] J.L. Pascual-ahuir, E. Silla, I. Tuñon, J. Comput. Chem. 15 (1994) 1127. (1993) 2968.
[60] J. Tomasi, WIREs: Comput. Mol. Sci. 1 (2011) 855. [104] R. Kirmse, J. Stach, W. Dietzsch, G. Steimecke, E. Hoyer, Inorg. Chem. 19 (1980)
[61] J. Tomasi, B. Mennucci, R. Cammi, Chem. Rev. 105 (2005) 2999. 2679.
[62] B. Mennucci, J. Tomasi, J. Chem. Phys. 106 (1997) 5151. [105] K.W. Plumlee, B.M. Hoffman, J.A. Ibers, Z.G. Soos, J. Chem. Phys. 63 (1975)
[63] M. Cossi, V. Barone, R. Cammi, J. Tomasi, Chem. Phys. Lett. 255 (1996) 327. 1926.
[64] D. Sellmann, S. Funfgelder, F. Knoch, M. Moll, Z. Naturforsch. (B) 46 (1991) [106] R. Kirmse, J. Stach, W. Dietzsch, E. Hoyer, Inorg. Chim. Acta 26 (1978) L53.
1601. [107] D. Snaathorst, H.M. Doesburg, J. Perenboom, C.P. Keijzers, Inorg. Chem. 20
[65] M. Fourmigue, J.N. Bertran, Chem. Commun. (2000) 2111. (1981) 2526.
[66] C. Mahadevan, M. Seshasayee, A. Radha, P.T. Manoharan, Acta Crystallogr. [108] J. Stach, R. Kirmse, W. Dietzsch, R.M. Olk, E. Hoyer, Inorg. Chem. 23 (1984)
Sect. C: Cryst. Struct. Commun. 40 (1984) 2032. 4779.
[67] P. Kuppusamy, P.T. Manoharan, J. Indian Chem. Soc. 63 (1986) 95. [109] E.J. Reijerse, A.H. Thiers, R. Kanters, M.C.M. Gribnau, C.P. Keijzers, Inorg. Chem.
[68] A. Husain, S.A.A. Nami, S.P. Singh, M. Oves, K.S. Siddiqi, Polyhedron 30 (2011) 26 (1987) 2764.
33. [110] R. Kirmse, B.V. Solovev, J. Inorg. Nucl. Chem. 39 (1977) 41.
[69] V.N. Staroverov, G.E. Scuseria, J. Tao, J.P. Perdew, Phys. Rev. B 69 (2004). [111] B.D. Towey, R.K. Szilagyi, Inorg. Chem., manuscript in preparation.
[70] J.P. Perdew, A. Ruzsinszky, L.A. Constantin, J. Sun, G.I. Csonka, J. Chem. Theory [112] T. Sriskandakumar, H. Petzold, P.C.A. Bruijnincx, A. Habtemariam, P.J. Sadler,
Comput. 5 (2009) 902. P. Kennepohl, J. Am. Chem. Soc. 131 (2009) 13355.
[71] C.T. Lee, W.T. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785. [113] A. Karunakaran-Datt, P. Kennepohl, J. Am. Chem. Soc. 131 (2009) 3577.
[72] J.P. Perdew, Phys. Rev. B 33 (1986) 8822. [114] V. Martin-Diaconescu, P. Kennepohl, Inorg. Chem. 48 (2009) 1038.
[73] C. Adamo, V. Barone, J. Chem. Phys. 108 (1998) 664. [115] N. Bohr, Philos. Mag. 26 (1913) 1.
[74] K. Molawi, A.J. Cohen, N.C. Handy, Int. J. Quantum Chem. 89 (2002) 86. [116] H.G.J. Moseley, Philos. Mag. 26 (1913) 1024.
[75] W.M. Hoe, A.J. Cohen, N.C. Handy, Chem. Phys. Lett. 341 (2001) 319. [117] P.R. Sarode, S. Ramasesha, W.H. Madhusudan, C.N.R. Rao, J. Phys. C: Solid State
[76] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865. Phys. 12 (1979) 2439.
[77] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C. [118] A. Manthiram, P.R. Sarode, W.H. Madhusudan, J. Gopalakrishnan, C.N.R. Rao,
Fiolhais, Phys. Rev. B 46 (1992) 6671. J. Phys. Chem. 84 (1980) 2200.
[78] J.P. Perdew, K. Burke, Y. Wang, Phys. Rev. B 54 (1996) 16533. [119] P.V. Khadikar, N.F. Mangelson, S.P. Pandharkar, Jpn. J. Appl. Phys. Part 1: Reg.
[79] Y. Zhao, D.G. Truhlar, J. Chem. Phys. 125 (2006). Pap. Short Notes Rev. Pap. 28 (1989) 709.
[80] J.M. Tao, J.P. Perdew, V.N. Staroverov, G.E. Scuseria, Phys. Rev. Lett. 91 (2003). [120] R.K. Katare, S.K. Joshi, B.D. Shrivastava, K.B. Pandeya, A. Mishra, X-ray Spec-
[81] E.E. Dahlke, D.G. Truhlar, J. Phys. Chem. B 109 (2005) 15677. trom. 29 (2000) 187.
[82] A.D. Becke, J. Chem. Phys. 104 (1996) 1040. [121] L. El Khoury, L. Journel, R. Guillemin, S. Carniato, W.C. Stolte, T. Marin, D.W.
[83] A.D. Becke, J. Chem. Phys. 98 (1993) 5648. Lindle, M. Simon, J. Chem. Phys. 136 (2012).
[84] A.D. Becke, J. Chem. Phys. 98 (1993) 1372.

You might also like