Sie sind auf Seite 1von 7

Inorganica Chimica Acta 485 (2019) 42–48

Contents lists available at ScienceDirect

Inorganica Chimica Acta


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

Research paper

Oxido-alcoholato/thiolato-molybdenum(VI) complexes with a dithiolene T


ligand generated by oxygen atom transfer to the molybdenum(IV)
complexes
Hideki Sugimotoa, , Masanori Satoa, Kaoru Asanob, Takeyuki Suzukib, Takashi Ogurac,

Shinobu Itoha,

a
Department of Material and Life Science, Division of Advanced Science and Biotechnology, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita,
Osaka 565-0871, Japan
b
Comprehensive Analysis Center, The Institute of Scientific and Industrial Research (ISIR), Osaka University, 8-1 Mihogaoka, Ibaraki, Osaka 567-0057, Japan
c
Picobiology Institute, Graduate School of Life Science, University of Hyogo, RSC-UH Leading Program Center, 1-1-1 Koto, Sayo-cho, Sayo-gun, Hyogo 678-0057, Japan

ARTICLE INFO ABSTRACT

Keywords: Oxido-alcoholato- and oxido-thiolato-molybdenum(VI) complexes bearing two ene-1,2-dithiolate ligands (cy-
Molybdenum(IV) clohexene-1,2-dithiolate) are prepared as synthetic models of molybdenum(VI) reaction centers of dimethyl
Molybdenum(VI) sulfoxide reductase family of molybdenum enzymes. These complexes are prepared by oxygen atom transfer
Dithiolene from tertiary amine N-oxide (trimethylamine N-oxide and N,N-dimethylaniline N-oxide) to the five-coordinate
Enzyme model
alcoholato- and thiolato-molybdenum(IV) complexes, and are characterized by UV–vis, cold-spray-ionization
Oxygen atom transfer
mass, resonance Raman, and 1H NMR spectroscopies. The oxygen atom transfer reactions are studied kinetically
at a low temperature (−40 °C) to demonstrate that the reactivity of the thiolato-molybdenum(IV) complex is
higher than that of alcoholato-molybdenum(IV) complex by about 7 times, and that the oxygen atom transfer
reactivity increases with increasing the electron withdrawing ability of the p-substituent of N,N-dimethylaniline
N-oxide derivatives. Mechanistic details are discussed based on the reactivity studies.

1. Introduction induce OAT from nitrate (NO3−, nitrate reductase) [1,2,4,5,6] Since the
oxygen atom abstraction becomes more difficult in the order of
The majority of molybdenum enzymes catalyzes net oxygen atom Me3NO < DMSO < NO3−, the molybdenum centers with −O2C·Asp
transfer (OAT) reactions from water to the substrates (H2O + S → and −S·Cys residues may have higher reactivity as compared to that
SeO + 2e− + 2H+) or their backward reactions [1–3]. The enzymes with −O·Ser.
constitute the oxotransferase class of molybdenum enzymes, most of In the modeling studies of the active sites of the DMSOR family
which structurally belong to dimethylsulfoxide reductase (DMSOR) molybdenum enzymes, molybdenum complexes formulated as MoIV(X)
family [1–3]. The OAT reactions occur between the molybdenum(IV) (ene-1,2-dithiolate)2 and MoVIO(X)(ene-1,2-dithiolate)2 have been re-
and molybdenum(VI) centers of the enzymes [1,2]. The DMSOR family vealed to reproduce stereo-chemical and electronic features related to
is featured by the presence of a mononuclear molybdenum center co- the coordination environments of the enzyme active sites [7–14]. Holm
ordinated by two ene-1,2-dithiolate chelate ligands, that is, the pyr- and co-workers have developed the alcoholato-, thiolato, and carbox-
anopterin cofactors (MPT) [1]. The molybdenum(IV) centers exhibit a ylate-molybdenum(IV) centers with bis(ene-1,2-dithiolate) chelates,
five-coordinate structure with one amino acid residue (X), whereas the using 1,2-dimethylethylene-1,2-dithiolate (S2C2Me2) or 1,2-dipheny-
molybdenum(VI) centers have one additional oxido group (Fig. 1), that lethylene-1,2-dithiolate (S2C2Ph2), as the model complexes of the re-
comes from the oxygenated substrates (SeO) [1–3]. The enzymes duced states of the DMSOR family [7,15–19]. With respect to the che-
having a molybdenum(IV) center coordinated by −O·Ser catalyze OTA mical functions, however, only the reactivity of the alcoholato-
reactions from DMSO and trimethylamine N-oxide (Me3NO, trimethy- moybdenum(IV) complexes was examined in the OAT reactions from
lamine N-oxide reductase), while the enzymes possessing a mo- Me3NO and DMSO, because the reactions of the thiolato- and carbox-
lybdenum(IV) center coordinated by −O2C·Asp or −S·Cys residue ylato-molybdenum(IV) complexes with those substrates (SeO) were too


Corresponding authors.
E-mail addresses: sugimoto@mls.eng.osaka-u.ac.jp (H. Sugimoto), shinobu@mls.eng.osaka-u.ac.jp (S. Itoh).

https://doi.org/10.1016/j.ica.2018.10.001
Received 25 July 2018; Received in revised form 1 October 2018; Accepted 1 October 2018
Available online 03 October 2018
0020-1693/ © 2018 Elsevier B.V. All rights reserved.
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

chloroperbenzoic acid (m-CPBA) [23]. Mo(CO)2L2 was synthesized by


following the established procedure, where L is cyclohexene-1,2-di-
thiolate [24].

2.2. Synthesis of [(Et4N)[MoIV(OEt)L2] (MoIV(OEt))

To a THF solution (2 mL) of Mo(CO)2L2 (44.9 mg, 101.9 µmol), a


1 mL acetonitrile solution of LiOEt (6.0 mg, 115.4 µmol) was added.
The mixture was stirred for 2 h. The color of purple solution was
X = O•Ser, S•CysS, O2C•Asp
changed to brown. To the brown solution, Et4NCl (17.4 mg,
Fig. 1. The active site structures of DMSOR family of molybdenum enzymes. 105.0 µmol) was added, and the solution was stirred for 10 min. After
filtration, the brown filtrate was concentrated to dryness to give a black
complicated to follow. Thus, the expected products, oxido-molybdenum material, which was recrystallized from acetonitrile/ether to give
(VI) complexes, [MoVIO(X)(S2C2R2)2]− (R = Me and Ph), were not brown crystals. Yield 36.3 mg (65%). Anal. Calcd. for
characterized [15–19]. Therefore, effects of the ligand X at the axial C22H41MoNOS4·0.5H2O (MW = 568.78): C 46.46; H 7.44; N 2.46.
position in the MoIV(X)(ene-1,2-dithiolate)2 complexes on the OTA re- Found: C 46.58; H 7.75; N 2.67%. 1H NMR (CD3CN, anionic part): δ
actions have not been clarified yet. 0.84 (t, 3H), 1.70–1.89 (m, 8H), 2.68–2.86 (m, 8H), 3.53 (q, 2H).
Recently, we have succeeded to prepare the series of bis(ene-1,2- UV–vis (MeCN, 25 °C): λmax = 326 nm (ε = 10000 M−1 cm−1), 378
dithiolate)molybdenum(VI)(O)(O2CR/OR/SR) complexes at low tem- (sh), 447 (sh).
perature as the models of oxidized-DMSOR [20–22]. The oxido-car-
boxylato-molybdenum(VI) complexes were prepared by an addition of 2.3. Synthesis of [(Et4N)[MoIV(OPh)L2] (MoIV(OPh))
the carboxylate anion to the five-coordinate oxido-molybdenum(VI)
complexes, and the oxido-alcoholato- and oxido-thiolato-molybdenum An acetonitrile suspension (2 mL) containing of Mo(CO)2L2
(VI) complexes were synthesized by substitution of one of the one oxide (47.2 mg, 107.2 µmol) and NaOPh (13.7 mg, 118.0 µmol) was stirred
groups of dioxide-molybdenum(VI) complex with alcohol or thiol in the for 3 h. The suspension became a purple solution and then the color of
presence of an acid. However, direct formation of the molybdenum(VI)- purple solution was changed to brown. To the brown solution, Et4NCl
oxido complexes from the corresponding molybdenum(IV) complexes (18.1 mg, 109.2 µmol) was added, and the solution was stirred for
by the OAT reaction with the enzymatic substrates (SeO) has yet to be 5 min. After filtration, the filtrate was concentrated to dryness to give a
accomplished. black solid. Yield 28.5 mg (44%). Anal. Calcd. for C26H41MoNOS4
In this study, we have examined the effects of the coordinating atom (MW = 607.82): C 51.38; H, 6.80; N, 2.30. Found: C 51.44; H 6.89; N
X on the OAT reactions of bis(ene-1,2-dithiolate)MoIV(OR/SR) model 2.39%. 1H NMR (CD3CN, anionic part): δ 1.73–1.91 (m, 8H), 2.77–2.94
complexes, and characterized the generated MoVIO(OR) and MoVIO(SR) (m, 8H), 6.36 (d, 2H), 6.76 (dd, 1H), 6.98 (t, 2H). UV–vis (MeCN,
product complexes. The complex structures and their abbreviations are 25 °C): λmax = 351 nm (ε = 14300 M−1 cm−1), 398 (4700), 473
given in Fig. 2. (2000).

2.4. Synthesis of [(Et4N)[MoIV(OPr2Ph)L2] (MoIV(OPr2Ph))


2. Experimental section
To a THF (1 mL) solution of Mo(CO)2L2 (44.9 mg, 101.9 µmol), a
2.1. General procedures 1 mL acetonitrile solution of Li(OPr2Ph) (19.0 mg, 103.1 µmol) was
added. The mixture was stirred for 16 h. The color of purple solution
All reagents and solvents were used as received unless otherwise was changed to brown. To the brown solution, Et4NCl (18.0 mg,
noted. All reactions were carried out under argon in a Miwa DB0-1KP 108.6 µmol) was added, and the solution was stirred for 10 min. After
glovebox. Propionitrile was dried with CaH2 and then P2O5 and distilled filtration, the filtrate was concentrated to dryness to give a black solid.
under dinitrogen atmosphere prior to use. All reactions were carried out Yield 44.2 mg (63%). Anal. Calcd. for C32H53MoNOS4·H2O
under a dinitrogen or argon atmosphere using standard Schlenk tech- (MW = 710.00): C 54.13; H 7.81; N 1.97. Found: C 54.42; H 8.04; N
niques or a Miwa D80-1KP glovebox. The compounds NaOPh, LiOEt, Li 2.10%. 1H NMR (CD3CN, anionic part): δ 0.97 (d, 12H), 1.71 ∼ 1.87
(OPr2Ph) Li(SPr3Ph) were prepared from the corresponding phenol, al- (m, 8H), 2.71–2.94 (m, 8H), 3.15 (sep, 2H), 6.66–6.72 (m, 1H),
cohol, or thiol using Na or nBuLi (1.6 M in hexane). (p-X-C6H4)Me2NO 6.74–6.79 (m, 2H). UV–vis (MeCN, 25 °C): λmax = 343 nm
were prepared from the corresponding N,N–dimethylaniline using m- (ε = 14000 M−1 cm−1), 411 (5500), 470 (sh 2200).

2.5. Synthesis of [(Et4N)[MoIV(OPr2Ph)L2] (MoIV(SPr3Ph))

To an acetonitrile/THF (1:1) solution (2 mL) containing Mo(CO)2L2


(44.1 mg, 100.1 µmol) and Li(SPr3Ph) (24.8 mg, 105.3 µmol) was stirred
for overnight. The color of purple solution was changed to brown. To
the brown solution, an acetonitrile solution (1 mL) of Et4NCl (17.9 mg,
108.0 µmol) was added, and the solution was stirred for 10 min. After
filtration, the brown filtrate was concentrated to dryness to give a black
solid. The solid was recrystallized from THF/ether to give brown
crystals. Yield 33.3 mg (44%). Anal. Calcd. for C35H59MoNS5·0.5H2O
(MW = 759.13): C 55.38; H 7.97; N 1.85. Found: C 55.31; H 8.07; N
1.86%. 1H NMR (CD3CN, anionic part): δ 1.03 (d, 12H), 1.16 (d, 6H),
1.70–1.85 (m, 8H), 2.76 (sep, 1H), 2.92–3.03 (m, 8H), 3.55 (sep, 2H),
6.73 (s, 2H). UV–vis (MeCN, 25 °C): λmax = 331 nm
Fig. 2. ChemDraw structures of molybdenum complexes and their abbrevia- (ε = 9600 M−1 cm−1), 349 (9700), 423 (sh), 483 (sh), 553 (sh), 770
tions. (sh).

43
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

2.6. Physical measurements 2.9. X-ray crystallography

FT–IR spectra were recorded with a Jasco FT-IR 4100 spectrometer. Single crystals of MoIV(OEt), MoIV(OPh), MoIV(OPr2Ph), and −
1
H NMR spectra were recorded with a JEOL Lambda 300, and TMS were obtained by vapor diffusion of diethyl ether into the acetonitrile
signal was adjusted to 0 ppm. UV–vis spectra were recorded on a HP- solutions. Each single crystal obtained was mounted on a loop, and all
8453 equipped with a Unisoku thermostat cell holder (USP–203). CSI- measurements were made on a Rigaku RAXIS RAPID imaging plate area
MS (cold electrospray ionization mass spectrum) measurements were detector with graphite monochromated Mo-Kα radiation. The structures
performed on a BRUKER cryospray microTOFII. Resonance Raman were solved by direct methods (SIR2011) and refined anisotropically by
scattering was excited at 590 nm with an Ar + ion laser (Spectra full–matrix least squares on F2 [25]. The hydrogen atoms were attached
Physics, BeamLok 2080). Resonance Raman scattering was dispersed by at idealized positions on carbon atoms and were not refined. All
a 75-cm single spectrometer (SPEX750M; Jovin Yvon) and was detected structures in the final stages of refinement showed no movement in the
by a liquid nitrogen cooled CCD detector (Spec10:400B/LN; Roper atom positions. The calculations were performed using Single–Crystal
Scientific). The resonance Raman measurement was carried out using a Structure Analysis Software, version 3.8 [26]. Crystallographic para-
spinning NMR tube at −80 °C by flashing cooled dinitrogen gas. meters are summarized in Table 1.

3. Results and discussion


2.7. Electrochemistry
3.1. Synthesis, characterization, and crystal structures of MoIV(OEt),
Cyclic voltammetric measurements were performed under dini- MoIV(OPh), MoIV(OPr2Ph), and −
trogen atmosphere with a Hokuto Denko HZ-3000 potentiostat. A set of
a glassy-carbon working electrode (circular, 3 mm diameter), a SCE The cyclohexene-1,2-dithiolate ligand, L, was employed to synthe-
reference electrode, and a platinum counter electrode was employed in size the series of bis(ene-1,2-dithiolate)molybdenum(VI) complexes,
these experiments. since we have already reported the dioxido- (MoVIO2) and oxido-ben-
zoato-molybdenum(VI) complexes, (MoVIO(OBz)), using this dithio-
lene ligand [21,23]. According to the synthetic procedures reported by
2.8. Kinetic experiments Holm and his co-workers [16,17], the new molybdenum(IV) complexes
having alcoholate or thiolate ligand at an axial position were synthe-
Kinetic measurements for the reactions of MoIV(OR/SR) with tri- sized (see Fig. 1). Fig. 3a, b, and c show the crystal structures of the
methylamine N-oxide or N,N-dimethylaniline N-oxide or its para-sub- anionic part of MoIV(OEt), MoIVO(OPh), and MoIV(OPr2Ph), respec-
stituted derivatives were performed using a Hewlett Packard 8453 tively. The selected bond lengths and angles are given in Table 2. The
photodiode array spectrometer, with a magnetic stirrer, equipped with Mo1 atoms are ligated by one oxygen atom (O1) of the alcoholate li-
a Unisoku thermostated cell holder designed for low temperature gand and four sulfur atoms, S1eS4, of the two ene-1,2-dithiolate li-
measurements (USP-203, desired temperature can be fixed gands. The molybdenum centers exhibit a square-pyramidal geometry
within ± 0.5 °C) in acetonitrile. To an acetonitrile or propionitrile so- having the alcoholate or phenolate monodetate ligand at the apical
lution of the 0.1 mM complex in a 1.0 cm path length UV–vis cell closed position and the four sulfur atoms of L at the basal plane, where the
with a silicon rubber septum cap, an acetonitrile solution of the sub- S1eS2eS3eS4 torsion angles is 2.27–5.29°. The stereochemistry is ty-
strate was added under pseudo-first-order conditions using a micro- pical as a bis(ene-1,2-dithiolate)des-oxomolybdenum(IV) complex.
syringe through the silicon rubber septum cap. The progress of the re- Also, the dimensions (bond distances of the SeCeCeS units) of L in-
action was monitored by following a decrease of the absorption band dicate that the ligand L take the ene-1,2-dithiolate structure. The Mo1
near 350 nm due to MoIV(OR/SR). The pseudo-first-order rate con- atoms are moved toward the apical ligand atom from the basal plane by
stants for the reactions were calculated from the plots of ln(ΔA) vs. time 0.76 Å for MoIV(OEt), 0.76 Å for MoIV(OPh), and 0.80 Å for
based on the time course of the absorbance change. MoIV(OPr2Ph). The Mo1eO1eC13 bond angles of MoIV(OEt) and

Table 1
Crystallographic data for MoIV(OEt), MoIV(OPh), MoIV(OPr2Ph), and MoIV(SPr3Ph).
MoIV(OEt) MoIV(OPh) MoIV(OPr2Ph) MoIV(SPr3Ph)

formula C22H41MoNOS4 C26H41MoNOS4 C32H53MoNOS4 C35H59MoNO1S5


fw 559.75 607.80 691.96 766.10
size (mm) 0.25 × 0.21 × 0.10 0.44 × 0.42 × 0.11 0.05 × 0.01 × 0.01 0.41 × 0.28 × 0.05
temperature (K) 103 103 103 103
crystal system orthorhombic monoclinic monoclinic monoclinic
space group Pna21 (#33) C2/c (#15) P21/c (#14) C2/c (#15)
Z 4 8 4 8
a, Å 9.4106(5) 25.7243(19) 9.2064(7) 36.086(3)
b, Å 11.9997(8) 8.0684(5) 16.8332(13) 16.0557(12)
c, Å 23.3106(16) 28.1414(15) 22.53978(14) 16.0079(14)
α, deg 90 90 90 90
β, deg 90 105.378(2) 99.895(3) 108.050(2)
γ, deg 90 90 90 90
V, Å3 2632.3(5) 5631.8(6) 3441.1(4) 8818.3(13)
μ, cm−1 8.285 7.810 6.480 5.6570
no. reflns (I > 2σIo) 5778 6442 7813 6347
no. reflns (all data) 24,737 25,917 32,006 10,098
no. variables 307 349 395 407
GOF 1.130 1.383 1.390 1.793
R1% (I > 2.0σI) 0.0430 0.0809 0.0838 0.1057
wR2% (all data) 0.1117 0.1920 0.2142 0.2610

R1 = Σ(|Fo|–|Fc|)/Σ|Fo|. wR2 = {Σ (w(Fo2–Fc2)2)/Σ w(Fo2)2}1/2.

44
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

(b) (c)
(a)
O1
O1
S4 S2 O1
C8 S4 S2
C2 C8 S4 S2
C2
Mo1 C8 C2
Mo1
C7 Mo1
S3 C7 S3
S1 C1 S1 C1 C7 S3 S1 C1

IV IV IV Pr2
Fig. 3. Crystal structures of the anionic parts of (a) Mo (OEt), (b) Mo (OPh), and (c) Mo (O Ph). Hydrogen atoms are omitted for clarity.

Table 2 Table 3
Selected bond lengths (Å) and angles (°) of MoIV(OR) complexes. Selected bond lengths (Å) and angles (°) of MoIV(SPr3Ph).
MoIV(OEt) MoIV(OPh) MoIV(OPr2Ph) MoIV(SPr3Ph)

Mo(1)eS(1) 2.3200(13) 2.3100(12) 2.3489(18) Mo(1)eS(1) 2.307(2) Mo(1)eS(4) 2.327(2)


Mo(1)eS(2) 2.3291(14) 2.3173(12) 2.3375(18) Mo(1)eS(2) 2.310(2) Mo(1)eS(5) 2.3318(19)
Mo(1)eS(3) 2.3342(10) 2.3389(12) 2.3233(19) Mo(1)eS(3) 2.3306(16)
Mo(1)eS(4) 2.3515(10) 2.3371(12) 2.3297(17)
S(1)eMo(1)eS(2) 83.09(8) S(2)eMo(1)eS(4) 86.79(7)
Mo(1)eO(1) 1.857(4) 1.899(2) 1.850(5)
S(1)eMo(1)eS(3) 86.64(7) S(2)eMo(1)eS(5) 106.73(7)
S(1)eMo(1)eS(2) 82.68(5) 82.37(4) 82.27(6) S(1)eMo(1)eS(4) 147.44(8) S(3)eMo(1)eS(4) 82.65(7)
S(1)eMo(1)eS(3) 83.75(4) 85.64(4) 84.14(7) S(1)eMo(1)eS(5) 104.67(8) S(3)eMo(1)eS(5) 111.03(6)
S(1)eMo(1)eS(4) 143.34(6) 139.87(4) 142.58(7) S(2)eMo(1)eS(3) 142.23(7) S(4)eMo(1)eS(5) 107.88(7)
S(1)eMo(1)eO(1) 107.26(12) 109.06(9) 108.87(14)
S(2)eMo(1)eS(3) 140.88(6) 143.11(3) 137.41(7)
S(2)eMo(1)eS(4) 86.77(4) 84.51(4) 84.30(6)
S(2)eMo(1)eO(1) 110.61(12) 109.42(9) 111.89(14) 3.2. Preparation and characterization of oxido-alcoholato- and oxido-
S(3)eMo(1)eS(4) 82.66(4) 82.55(4) 82.53(6)
thiolato-molybdenum(VI) complexes
S(3)eMo(1)eO(1) 108.44(12) 107.47(9) 110.69(15)
S(4)eMo(1)eO(1) 109.32(12) 111.07(9) 108.54(13)
We have previously reported that the di-oxido-molybdenum(VI)
complex of L was produced by the reaction of the 5-coordinate oxido-
molybdenum(IV) complex of L with trimethylamine N-oxide (Me3NO)
MoIV(OPh) are close to each other (133.3(4) ° and 135.4(2)°), whereas [23]. In this study, we have examined the reaction of the five-co-
that of MoIV(OPr2Ph) (160.5(4)°) is larger due to the steric hindrance of ordinate-alcoholato/thiolato-molybdenum(IV) complexes with Me3NO
the two isopropyl groups at the o-positions of OPr2Ph axial ligand. to see whether oxido-alcoholato- and oxido-thiolato-molybdenum(VI)
The crystal structure of MoIV(SPr3Ph) is shown in Fig. 4 together complexes are formed. In the reaction at room temperature, 5-co-
with the bond lengths and angles given in Table 3. The complex also ordinate mono-oxide molybdenum(V) complex was formed. This ob-
exhibits a square-pyramidal geometry as in the case of MoIV(OR), and L servation was consistent with the reaction of [MoIV(OBz)L2]− with
ligands take the ene-1,2-dithiolate structure as revealed by the bond Me3NO at room temperature. On the other hand, at low temperature,
distances of the SeCeCeS units. The distance between Mo1 and the we observed formation of a different species from the mono-oxide
basal plane defined by the four sulfur atoms from the two L ligands is molybdenum(V) complex. Fig. 5a illustrates the UV–vis spectral change
0.70 Å, which is shorter than those of the three alcoholate complexes of MoIV(OEt) upon an addition of Me3NO in CH3CN at −40 °C, where
(0.76–0.80 Å). Furthermore, the Mo1eSeC (monodentate thiolate) the absorption band at 324 nm due to MoIV(OEt) decreased, whereas
angle of 107.2(3)° is significantly smaller than those of the alcoholate those at 573 (3600) and 778 (ε = 2200 M−1 cm−1) nm increased as the
complexes (133.3(4)° for MoIV(OEt), 135.4(2)° for MoIV(OPh), reaction proceeded. The generated complex was stable below −40 °C,
160.5(4) for MoIV(OPr2Ph)) in spite of the presence of the two iso- and its spectrum was very similar to those of the oxido-(alcoholato/
propyl groups at the o-positions. This is due to a larger contribution of silanolato/phenolato)-molybdenum(VI) complexes, [MoVIO(OSiPh2iPr/
the p-orbitals of the S atom as compared with that of the p-orbitals of OEt)(S2C2(COOMe)2)2]− and [MoVIO(OBz)L2]−, reported previously
the O atom in the MoIV(OR) complexes. [20,22,21]. The CSI–mass spectrum of an acetonitrile solution of
MoIV(OEt) in the presence of a slightly excess amount of Me3NO ex-
hibited a peak cluster at m/z = 447.0 together with those at m/
z = 431.0 and 402.0 at −60 °C as shown in Fig. 5b. The peak cluster at
m/z = 447.0 is consistent with the molecular formula of the oxido-
ethanolato-molybdenum(VI) complex, [MoVIO(OEt)L2]−. The peak
clusters at m/z = 431.0 and 402.0 correspond to the fragments of
{[MoVIO(OEt)L2]–O}− and {[MoVIO(OEt)L2]–OEt}−, respectively. The
1
H NMR spectrum of an acetonitrile-d3 solution containing MoIV(OEt)
and Me3NO taken at −60 °C showed broad signals at 2.74 and
1.74 ppm due to the protons of the cyclohexene units of L and one
quartet and one triplet of the ethanolate ligand at 3.49 and 0.80 ppm.
These proton signals shifted upon the treatment with Me3NO to 2.47,
1.64, 3.71, and 1.25 ppm, respectively, together with appearance of
new singlet peak at 2.08 ppm due to Me3N as shown in Fig. S1. In ad-
Fig. 4. Crystal structure of the anionic part of MoIV(SPr3Ph). Hydrogen atoms
dition, the resonance Raman spectrum of a propionitrile solution
are omitted for clarity.

45
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

(a) (b)

1.5 324
Sim.

Absorbance
1.0

Exp. 447.0 [MoVIO(OEt)L2]-


573
445.0
444.0
446.0

0.5
449.0

778
441.0
443.0

448.0

442.0 450.0

440 445 450 455


0.0 m/z
400 600 800 1000
Wavelength / nm
402.0
447.0
431.0

400 450 500


m/z
Fig. 5. (a) UV–vis spectrum of MoVIO(OEt (0.10 mM, blue) generated from MoIV(OEt) (black) and Me3NO (5.0 mM) at −40 °C and (b) CSI-mass spectrum of MoVIO
(OEt) at −60 °C. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

containing MoIV(OEt) and Me3NO exhibited a Raman band at due to MoVIO(OEt) with clear isosbestic points at 307 and 401 nm. The
872 cm−1 that is assignable to a ν(Mo]O) stretching vibration (Fig. decay of the absorption band at 324 nm obeyed first-order kinetics as
S2). Weak intensity of the peak at 872 cm−1 may be due to instability of shown in the inset of Fig. 6a. The apparent first-order rate constant kobs
the generated complexes under photo-irradiation conditions. On the exhibited first-order dependence on the concentration of Me3NO
basis of these spectroscopic data (UV–vis, CSI-mass, 1H NMR, and (Fig. 6b). Thus, the oxygen atom transfer reaction can be expressed by
Raman spectra), we concluded that oxido-ethanolato-molybdenum(VI) the second-order rate law, v = k2[MoIV(OEt)][Me3NO], for which the
complex of L, MoVIO(OEt), was generated by oxygen atom transfer second-order rate constant k2 was calculated to be 5.0 × 10−2 M−1 s−1
(OAT) from Me3NO to MoIV(OEt). This is the first example of spec- from the slope of the linear plot of Fig. 6a. The kinetic analysis was also
troscopically characterized bis(ene-1,2-dithiolato)oxido-alcoholato- carried out the for the OAT reactions from Me3NO to the phenolato-,
molybdenum(VI) complex that was formed by OAT from an enzymatic 2,6-di-isopropyl-phenolato-, and 2,4,6-tri-isopropyl-thiophenolato-mo-
substrate (amine-oxido) to the des-oxido-molybdenum(IV) complex. lybdenum(IV) complexes. The spectroscopic changes and the kinetic
Under similar reaction conditions for MoIV(OEt), MoIV(OPh)L, and analyses are given in Figs. S10, S9, and S12, and the k2 values thus
MoIV(OPr2Ph) were treated with Me3NO in CH3CN or in C2H5CN at the obtained are listed in Table 4.
low temperature (−40 °C). Monitoring the reactions by the UV–vis, 1H Among the three alcoholato complexes, MoIV(OPh) exhibited the
NMR, CSI-mass, and resonance Raman spectroscopic methods revealed highest reactivity. This can be explained, if the less OPh− → MoIV
that the corresponding oxido-molybdenum(VI) complexes, MoVIO electron donation makes the MoIV center more Lewis acidic than that in
(OPh)L and MoVIO(OPr2Ph), were formed by the OAT reaction from MoIV(OEt), resulting in facile coordination of Me3NO to the MoIV
Me3NO (Figs. S3, S4, S5, and S6). The ν(Mo]O) stretches of MoVIO center, and if the two isopropyl groups at the 2 and 6 positions of
(OEt) at 872 cm−1 and of MoVIO(OPr2Ph) at 871 cm−1 are very close to MoIV(OPr2Ph) sterically prevent the coordination of Me3NO to the MoIV
that of Rh. Sphaeroides DMSOR (862 cm−1) [27]. center.
In a similar way, the oxido-thiolato-molybdenum(VI) complex In the case of MoIV(SPr3Ph), the reaction at −40 °C was too fast to
MoVIO(SPr3Ph) was generated by the OAT reaction from Me3NO to be followed by UV–vis spectroscopy. Thus, the kinetic study was carried
MoIV(SPr3Ph) but at a lower temperature (−60 °C). The UV–vis, CSI- out at −60 °C (Fig. S12). Then, the oxygen atom transfer reactivity of
mass, and resonance Raman spectra are given in Figs. S7, S8, and S9. MoIV(SPr3Ph) at −40 °C was estimated as 17.4 × 10−2 (M−1 s−1) by
MoVIO(SPr3Ph) showed a significant red shift of the absorption bands in extrapolating the reaction rate constants obtained at the lower tem-
the visible region (λmax (ε) = 640 (5100) and 830 nm perature to that at −40 °C in the Eyring plot shown in Fig. S13. Ap-
(1200 M−1 cm−1)) when compared with those of the alcoholate com- parently, the reactivity of MoIV(SPr3Ph) is 7 times higher than that of
plexes, MoVIO(OEt) (λmax (ε) = 573 (3600) and 778 nm MoIV(OPr2Ph). As revealed by the crystal structural analysis, the
(2200 M−1 cm−1)), MoVIO(OPh) (615 (3600) and 815 nm MoeSeC bond angle of 107.2(3)° in MoIV(SPr3Ph) is more acute when
(2100 M−1 cm−1), and MoVIO(OPr2Ph) (624 (3900) and 820 nm compared with the MoeOeC bond angle of 160.5(4)° in MoIV(OPr2Ph).
(2000 M−1 cm−1). Thus, MoIV(SPr3Ph) may offer a less crowded vacant site around the
molybdenum center, making the access of Me3NO to the MoIV center
3.3. Kinetic study of oxygen atom transfer reactions from tertiary amine N- easier. The significant contribution of p orbital character of the S atom
oxide to alcoholato-and thiolato-molybdenum(IV) complexes to the MoeSeC bond angle, as reflected by the bond angle that is close
to 90°, may have an electronic component that favors substrate binding.
An oxygen atom transfer (OAT) from Me3NO to MoIV(OR/SR) To get further insight into the reaction mechanism, N,N-dimethy-
complexes was then examined kinetically. Fig. 6a shows the UV–vis laniline N-oxide and its para-substituted derivatives ((p-A-C6H4)N
spectral change for the OAT reaction from Me3NO to MoIV(OEt) in (Me)2O) were employed as the substrates for the OAT reactions with
CH3CN at −40 °C. The absorption band at 324 nm due to MoIV(OEt) MoIV(OPr2Ph) and MoIV(SPr3Ph). Fig. S14 shows a typical example of
decreased with simultaneous increase of the bands at 573 and 778 nm the UV–vis spectral change for the OAT reaction from (C6H5)N(Me)2O

46
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

(a) (b)
0 1.5
324 nm
1.5
324

ln(A∞–A)
k2 = 5.0 x 10-2 M-1 s-1

Absorbance
1.0 1.0

103 kobs / s–1


-4
0 5000 10000
0.5
573 Time / s
778 0.5

0.0
400 600 800 1000
Wavelength / nm
0.0
0 5 10 15 20 25
103 [Me3NO] / M

Fig. 6. (a) UV–vis spectral changes for the OAT reaction from Me3NO (5.0 mM) to MoIVO(OEt) (0.10 mM) in CH3CN at −40 °C; inset, a first-order plot based on the
absorption change at 324 nm. (b) Plot of kobs against [Me3NO].

Table 4 substituents may accelerate the dissociation of the tertiary amine from
The k2 values for the OAT reactions from Me3NO to MoIV(OR) and MoIV(SR) at intermediate A (kd process), whereas the electron-donating substituents
−40 °C. may accelerate the addition of N-oxide to the Mo(IV) center (ka pro-
MoIV(OEt) MoIV(OPh) MoIV(OPr2Ph) MoIV(SPr3Ph) cess). The observed positive Hammettρvalues (Fig. 7) suggest that the ka
process is more dominant as compeered to the kd process in the present
102k2/M−1 s−1 5.0 320 2.5 17.4a OAT reaction. Moreover, the larger ρ value obtained with MoIV(SPr3Ph)
a suggests that the kd process becomes more significant in the oxygen
Extrapolated value using the Eyring plot shown in Fig. S13.
atom abstraction by MoIV(SPr3Ph).
to MoIV(OPr2Ph) in CH3CN at −40 °C. The absorption band at 340 nm
due to MoIV(OPr2Ph) decreased with simultaneous increase of the 4. Conclusion
bands at 624 and 820 nm due to MoVIO(OPr2Ph), keeping clear iso-
sbestic points at 316, 396, and 426 nm. The λmax and ε values of the In this study, we have prepared a series of bis(ene-1,2-dithiolate)-
final spectrum were identical to those of MoVIO(OPr2Ph) characterized oxido-alcoholato- and oxido-thiolato-molybdenum(VI) complexes as
above. Then, the rate constants were plotted against the substituent the models for oxidized DMSOR family of molybdenum enzymes by the
constants σ to give a relatively small positive Hammett ρvalues in both OAT reaction from tertiary amine N-oxides to des-oxido molybdenum
cases (ρ = +0.52 for MoIV(OPr2Ph) and ρ = +0.80 for MoIV(SPr3Ph)). (IV) complexes. The molybdenum(VI) complexes were stable enough to
The results clearly indicate that the OAT reactivity increases as the be characterized at low temperature. Kinetic study for the OAT reaction
electron density of the oxygen atom of N-oxide becomes smaller. Such revealed that the reactivity for the OAT by MoIV(SR) is 7 times higher
results can be explained by the reaction involving two steps; (i) re- than that by MoIV(OR). Such a higher reactivity of MoIV(SR) can be
versible addition of the tertiary amine N-oxide to the molybdenum(IV) attributed in part to the less crowded coordination environment around
center of MoIV(OPr2Ph) and MoIV(SPr3Ph) generating an intermediate the molybdenum center as revealed by the acute MoeSeC bond angle.
A (ka and k−a) and (ii) dissociation of the tertiary amine from inter- The present results will provide important insights into the catalytic
mediate A (kd) as illustrated in Fig. 8. If so, the electronic effect of the p- mechanism as well as spectroscopic properties of the reaction centers in
substituents may operate oppositely. Namely, the electron withdrawing the molybdenum enzymes.

(a) (b)
1.0 1.0
MoIV(OPr2Ph) MoIV(SPr3Ph)

CN
log(k2(A) / k2(H))
log(k2(A) / k2(H))

0.5 0.5 Br
Br
CN
Me
0.0 OMe 0.0
OMe
H H
Me ρ = +0.52 ρ = +0.80
-0.5 -0.5
-0.8 -0.4 0.0 0.4 0.8 -0.8 -0.4 0.0 0.4 0.8
σ σ
Fig. 7. Hammett plots for the OAT reactions from (p-A-C6H4)N(Me)2O to (a) MoIV(OPr2Ph) in CH3CN at −40 °C and (b) MoIV(SPr3Ph) in C2H5CN at −60 °C.

47
H. Sugimoto et al. Inorganica Chimica Acta 485 (2019) 42–48

Fig. 8. Possible reaction mechanism for the OAT re-


action from (p-X-C6H4)N(Me)2O to the MoIV com-
plexes.

Acknowledgements 275 (1997) 1305.


[6] M. Jormakka, D. Richardson, B. Byrne, S. Iwata, Structure 12 (2004) 95.
[7] J.H. Enemark, J.J. Cooney, J.-J. Wang, R.H. Holm, Chem. Rev. 104 (2004) 1175.
This work was supported by a Grant-in-Aid for Scientific Research [8] J. McMaster, J.M. Tunney, C.D. Garner, Prog. Inorg. Chem. 52 (2004) 539.
on Innovative Areas “Stimuli-responsive Chemical Species for the [9] H. Sugimoto, H. Tsukube, Chem. Soc. Rev. 37 (2008) 2609.
Creation of Functional Molecules (No. 2408)” (JSPS KAKENHI Grant [10] F.J. Hine, A.J. Taylor, C.D. Garner, Coord. Chem. Rev. 254 (2010) 1570.
[11] C. Schulzke, Eur. J. Inorg. Chem. 2011 (2011) 1189.
Number JP24109015). [12] R.H. Holm, E.I. Solomon, A. Majumdar, A. Tenderholt, Coord. Chem. Rev. 255
(2011) 993.
Appendix A. Supplementary data [13] A. Majumdar, S. Sarkar, Coord. Chem. Rev. 255 (2011) 1039.
[14] A. Majumdar, Dalton Trans. 43 (2014) 8990.
[15] B.S. Lim, J.P. Donahue, R.H. Holm, Inorg. Chem. 39 (2000) 263.
Figs. S1–S14 showing UV-vis, 1H NMR, and resonance Raman [16] B.S. Lim, K.-M. Sung, R.H. Holm, J. Am. Chem. Soc. 122 (2000) 7410.
spectra as well as kinetic data. CCDC 1857754 and 1857756-1857758 [17] B.S. Lim, R.H. Holm, J. Am. Chem. Soc. 123 (2001) 1920.
[18] J.-J. Wang, O.P. Kryatova, E.V. Rybak-Akimova, R.H. Holm, Inorg. Chem. 43 (2004)
contain the supplementary crystallographic data for complexes
8092.
MoIV(OEt), MoIV(OPh), MoIV(OPr2Ph), and MoIV(SPr3Ph). These data [19] J.-J. Wang, C. Tessier, R.H. Holm, Inorg. Chem. 45 (2006) 2979.
can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/ [20] H. Sugimoto, S. Tatemoto, K. Suyama, H. Miyake, R.P. Mtei, S. Itoh, M.L. Inorg,
retrieving.html, or from the Cambridge Crystallographic Data Centre, Chem. 49 (2010) 5368.
[21] H. Sugimoto, M. Sato, L.J. Giles, K. Asano, T. Suzuki, M.L. Kirk, S. Itoh, Dalton.
12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336 033; or e- Trans. 42 (2013) 15927.
mail: deposit@ccdc.cam.ac.uk. Supplementary data to this article can [22] H. Sugimoto, M. Sato, K. Asano, T. Suzuki, K. Mieda, T. Ogura, T. Matsumoto,
be found online at https://doi.org/10.1016/j.ica.2018.10.001. L.J. Giles, A. Pokhrel, M.L. Kirk, S. Itoh, Inorg. Chem. 55 (2016) 1542.
[23] R.S. Lewis, M.F. Wisthoff, J. Grissmerson, W.J. Chain, Org. Lett. 16 (2014) 3832.
[24] H. Sugimoto, M. Harihara, M. Shiro, K. Sugimoto, K. Tanaka, H. Miyake,
References H. Tsukube, Inorg. Chem. 44 (2005) 6386.
[25] M.C. Burla, R. Caliandro, M. Camalli, B. Carrozzini, G.L. Cascarano, C. Giacovazzo,
M. Mallamo, A. Mazzone, G. Polidori, R. Spagna, J. Appl. Cryst. 45 (2012) 357.
[1] M.J. Romao, Dalton Trans. 38 (2009) 4053.
[26] Crystal Structure 3.8: Crystal Structure Analysis Package, Rigaku Corporation
[2] R. Hille, Dalton Trans. 42 (2013) 3029.
(2000-2006). 9009 New Trails Dr. The Woodlands TX 77381 USA.
[3] R. Hille, P. Basu, Chem. Rev. 114 (2014) 3963.
[27] S.D. Garton, J. Hilton, H. Oku, B.R. Crouse, K.J. Rajagopalan, M.K. Johnson, J. Am.
[4] C. Sparacino-Watkins, J.F. Stolz, P. Basu, Chem. Soc. Rev. 43 (2014) 676.
Chem. Soc. 119 (1997) 12960.
[5] J.C. Boyington, V.N. Gladyshev, S.V. Khangulov, T.C. Stadtman, P.D. Sun, Science

48

Das könnte Ihnen auch gefallen