You are on page 1of 6

DOI: 10.1002/slct.

201800163 Full Papers

1
2 z Catalysis
3
4
5
Synthesis and Structure of Ru(II) Complexes of
6 Thiosemicarbazone: Highly Selective Catalysts for
7
8 Oxidative Scission of Olefins to Aldehydes
9
10 Subramanian Muthumari and Rengan Ramesh*[a]
11
12
13 New Ru(II)-TSC complexes of the type [RuH(TSC)(CO)(PPh3)2] oxidative cleavage of a wide range of alkenes as well as alkynes
14 (where TSC = monobasic bidentate anthracenealdehyde thio- to the corresponding aldehydes/ diketones in acetonitrile/
15 semicarbazone ligands) were synthesized and characterized by water in the presence of NaIO4 at room temperature. The
16 means of elemental analysis, IR, UV vis, and NMR spectral maximum yield was obtained up to 99% using 0.5 mol% of
17 methods. Molecular structures of the complexes were deter- catalyst loading without any over oxidation byproducts. The
18 mined by single crystal X-ray crystallography, which prove the effect of substituents of the ligand, solvents, reaction temper-
19 coordination mode of the ligands and reveal the presence of a ature and catalyst loading on the catalytic activity of the
20 distorted octahedral geometry around the Ru ion. The synthe- complexes is also investigated.
21 sized complexes were developed as catalysts for selective
22
23 Introduction
conditions are still not resolved. Oxidation catalysis by
24
The process of oxidation of olefins to the subsequent carbonyl ruthenium-based complexes has received much interest due its
25
compounds, without over oxidation to carboxylic acid, is the well-known redox states of ruthenium.[5] Common side reac-
26
many useful method for fine chemical synthesis.[1] This versatile tions in the oxidation of olefinic double bonds such as
27
process transforms hydrocarbons to more valuable oxygenated epoxidation, dihydroxylation, and allylic oxidation reactions
28
derivatives and is also used in biomass processing and for the were avoided by using of catalytic amount of ruthenium and
29
synthesis of bioactive compounds thus rendering the process its complexes.
30
both industrially and academically relevant. Traditionally, Recently, Bera and co-workers have reported Ru(II) based
31
ozonolysis is one of the common methods to obtain aldehydes abnormal NHC complexes as highly active catalysts for
32
or carboxylic acids. Unfortunately, to achieve this researchers oxidative scission of olefins to aldehydes with 1 mol% up to
33
are still seriously limited by the use of the dangerous and 100% conversions within 30 min.[6] New (h5- cyclopentadienyl)
34
inconvenient ozonolysis method with a reductive workup, or dicarbonylruthenium(II) amine complexes (0.5 mol%) have
35
low selectivity and poor yielding Limieux- Johnson oxidation been reported as catalysts for oxidation of styrene in CH3CN/
36
protocol.[2] Recently, attention has been focused on transition H2O solvent system at 60 8C.[7] Catalytic performance of
37
metal catalysts in combination with environmentally benign ruthenium CNC pincer complex towards the oxidative cleavage
38
and clean oxidants for the development of sustainable of olefins with 1 mol% of catalyst loading for 24 h was explored
39
oxidation process. Among the various oxidants used, such as by Eduardo Peris and co-workers.[8] Dramatic effect of ancillary
40
NaIO4, H2O2, O2, oxone, PhI(OAc)2, and TBHP etc., NaIO4was NHC ligand present in ruthenium complex was found towards
41
found to be as an oxidant because of its solubility in water, highly selective catalytic oxidation of olefins/alkynes to
42
very cheap, easy handling procedure leads to broad applica- carbonyl compounds.[9] Environmentally benign solid catalyst
43
tions in nearly all the organic transformations.[3] To date, some RuO2/ BaTiO4 was used as a catalyst for the oxidative cleavage
44
reports on oxidation of olefins into aldehydes through Fe, Pd, of olefins to aldehyde and 3 mol % of catalyst and 6 equivalent
45
Os, Au Cu, Co catalysts,[4] biphasic PEG- 300/ SCCO2 catalyst and of NaIO4 oxidant in EtOAc– H2O buffer solution have been used
46
supported catalysts with NaIO4 been described. However, the in 3–21 h.[10] Chi-Ming Che et al. have documented ruthenium–
47
problems of low activity, low selectivity, environment un- porphyrin complexes as selective conversion of terminal
48
friendly, formation of side products and harsh reaction alkenes into corresponding aldehydes using 2 mol% catalyst in
49
presence of 2, 6-dichloropyridine N-oxide as oxidant.[11] Further,
50
[a] S. Muthumari, R. Ramesh Ronny Neumann and co-workers have reported [RuII(dmp)2
51
Centre for Organometallic Chemistry (H2O)2]2 + catalyst for the oxidative cleavage of olefins using
52 School of Chemistry H2O2 as the oxidant in CH3CN at 55 8C.[12] However, most of the
53 Bharathidasan University
reported oxidation reactions are suffer from high reaction
54 Tiruchirappalli 620 024, Tamil Nadu, India
E-mail: ramesh_bdu@yahoo.com temperature, long reaction time and high catalyst or oxidant
55
Supporting information for this article is available on the WWW under loadings and hence the development of new catalysts is
56
https://doi.org/10.1002/slct.201800163 essential.
57

ChemistrySelect 2018, 3, 3036 – 3041 3036  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3036/3041] 1
Full Papers
Particularly, metal complexes bearing thiosemicarbazone-
1
ligands have significant features of their ease of synthesis,
2
chelating capability, structural flexibility, as well as potential
3
biological activities.[13] The chemical interest arises from the
4
capability of thiosemicarbazone to adopt various coordination
5
modes, leading to enormous structural stability of their
6
complexes. It is noteworthy that TSC-based ruthenium catalysts
7
are not explored in carbon-carbon cleavage reactions. Follow-
8
ing to our previous works on the catalytic applications of
9
ruthenium and palladium complexes,[14] we wish to describe
10
herein the catalytic property of ruthenium(II) complex bearing
11 Scheme 2. The formation and structure of catalyst precursors [RuH
TSC as a spectator ligand in the selective oxidative scission of
12 (CO)(PPh3)2(TSC)].
olefins to aldehydes.
13
14
15
peared upon complexation. This observation may be attributed
16
to the enolisation of N-NH-C=S and subsequent coordination
17
through the deprotonated sulphur to the Ru(II)ion. The band
18
observed in the region 1313–1327 cm 1 due to nC-S, further
19
confirms the co-ordination through the deprotonated sulphur.
20
In addition, all the complexes show a intense band in the
21
region 1929–1951 cm 1 is due to the terminally coordinated
22
carbonyl group and is observed at a slightly higher frequency
23
than in the precursor ruthenium complexes. The other
24
characteristic bands due to coordinated PPh3 are also present
25
near 534, 698, 741, 1432–1438, 1481 cm 1.
26
The absorption spectra of the complexes were recorded in
27
acetonitrile at room temperature and showed three bands in
28
the region 223–390 nm. The high intensity bands in the region
29
223–322 nm were assigned to p-p* and n-p* transitions in the
30
ligand. In the complexes the lowest energy bands observed in
31
the region 383–390 nm were attributed to the Ru(dp)-to-L(p*)
32
metal-to-ligand charge transfer (MLCT) transitions. These prom-
33
inent strong bands of both intra ligand and charge transfer
34
absorptions support the presence of an octahedral environ-
35 Scheme 1. Selected reports for selective olefin oxidation reaction.
ment for the Ru metal ions, similar to that of other reported
36
octahedral ruthenium (II) complexes.[17]
37
In 1H NMR spectra, multiplets observed in the region d
38
8.38– 6.80 ppm in the complexes were due to the aromatic
39
protons of PPh3 and ligands. The appearance of a sharp singlet
40 Result and discussions
in the region of 9.17- 9.02 ppm has been assigned to
41
The TSC ligands HL1– HL3 were prepared from the reaction of azomethine proton (CH=N). The azomethine signal of the
42
anthracenealdehyde and thiosemicarbazides in methanol by ligands does not undergo any change even after the formation
43
literature method.[15] The new ruthenium TSC complexes (1-3) complexes clearly indicate that the azomethine nitrogen is not
44
were synthesized by reacting one equivalent of the anthrace- involved in the coordination. The N–NH–C=S proton of the free
45
nealdehyde thiosemicarbazone ligands with one equivalent of ligands shows a singlet at 10.08- 10.04 ppm which was absent
46
[RuHCl(CO)(PPh3)3][16] in benzene under reflux condition for 5 h in the complexes, supporting enolisation and coordination of
47
(Scheme 2). The complexes were obtained as brown solid in the thiolate sulfur to the ruthenium (II) ion. Further, Ru H
48
82–85% yield. The abstraction of a proton from the thiol protons were confirmed by 1H NMR spectra and were observed
49
sulphur was promoted by addition of triethylamine which at 12.08 to 11.73 ppm (Figures S1-S3 in supporting informa-
50
facilitate the coordination of the thiolate sulfur to the tion). Furthermore, the solid state molecular structure of the
51
ruthenium. All the complexes are soluble in CH2Cl2, CHCl3, complexes 1and 2 were studied by single crystal X-ray
52
CH3OH, DMF, DMSO, and CH3CN. diffraction (Figure 1 and 2) (Tables S1 and S2 in supporting
53
The coordination mode of the ligand to metal is confirmed information). Therefore, the ligand was coordinated to ruthe-
54
by selected bands of the IR spectra of the complexes 1 3. The nium ion through N(1) hydrazinic nitrogen and thiolate sulfur,
55
free ligand displayed nC = S and nN = H absorptions in the region forming a four-membered chelate ring. The formation of a four-
56
828–835 cm 1 and 3188–3199 cm 1 respectively which disap- membered ring may be due to presence of a bulky CO ligand
57

ChemistrySelect 2018, 3, 3036 – 3041 3037  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3037/3041] 1
Full Papers

1 Table 1. Influence of solvent, oxidant and temperature.[a]


2
3
4 Entry Solvent Oxidant Yield[b](%)
5
1 CH3CN/H2O NaIO4 97
6
2 CH3CN/H2O NaIO4 99c
7 3 CH3CN/H2O H2O2 85
8 4 CH3CN/H2O m-CPBA 82
t
9 5 CH3CN/H2O BuOOH 36
6 Acetone/H2O NaIO4 80
10
7 CH3CN/EtOAc/H2O(2:2:1) NaIO4 96
11 8 CH3CN/CCl4/H2O(2:2:1) NaIO4 89
12 9 EtOAc/H2O NaIO4 40
13 10 DCM/H2O NaIO4 34
11 H2O NaIO4 33
14
12 CH3CN NaIO4 65
15 13 CH3CN/H2O NaIO4 05d
16
[a] Conditions: 4-methylstyrene (1 mmol), catalyst 1 (1 mol %), oxidant
17 (2.0 mmol) in solvents (5 mL, 1:1) for 1 h at room temperature.
18 [b] Isolated yield.
19 Figure 1. ORTEP diagram of complex 1with thermal ellipsoids at the 50% [c] Reaction at 608 C.
probability level. [d] Reaction was performed in the absence of catalyst 1.
20
21
22
23
ties, the other two complexes are found to have similar
24
structure to that of [RuH(CO)(PPh3)2(L2)] (2). The X-ray determi-
25
nation reveals the structure expected on the basis of spectro-
26
scopic data, consistent with ruthenium ion in bivalency and the
27
mono ionized nature of the ligand in the complexes.
28
Literature studies show that there is no single experimental
29
procedure to get high yield in the catalytic oxidation of olefins.
30
Optimization of various oxidants, solvents (Table 1) and catalyst
31
loading (1 to 0.05 mol %, Table 2) were under taken using a test
32
33
34 Table 2. Optimization of catalyst loading.
35
36
37
Entry Mol% of Ru catalyst Yield (%) TON
38
39 1 1.0 99 99
2 0.5 96 192
40
Figure 2. ORTEP diagram of complex 2 with thermal ellipsoids at the 50% 3 0.3 81 270
41 probability level. 4 0.2 63 315
42 5 0.1 22 220
43 6 0.05 11 220
44 Conditions: 4-methylstyrene (0.2 mmol), oxidant (2.0 mmol) in solvents
or substitution on the azomethine carbon atom. Thus, the
45 CH3CN/H2O (5 mL, 1:1) for 1 h.
complex adopts a pseudo octahedral coordination environ-
46
ment around the central ruthenium (II) ion. Bond angles around
47
ruthenium(II) ion are N(1)–Ru(1)–S(1) = 64.43(6)8, P(1)–Ru(1)–
48
H(1) = 89.4(9)8, N(1)–Ru(1)–H(1) = 97.6(9) (Table 1). The bond reaction (1 mmol of 4-methylstyrene, 2 equivalent of oxidant,
49
lengths of Ru(1)–H(0), Ru(1)–N(1) and Ru(1)–S(4) are 1.86(3), 5 mL of solvent; Table 2). Screening of various oxidants
50
2.132(2) and 2.5663(8)Å, respectively. In the complex, the indicated that NaIO4 gave better result than other oxidants.
51
phosphine ligands are in trans conformation, with the P(1)–Ru– Quantitative yield of aldehyde was obtained in neutral
52
P(2) angle being equal to 173.65(3)8. Further, the Ru(1)– P(1) condition with 0.5 mol% of catalyst loading within 1 hour in
53
and Ru(1)– P(2) bond length of 2.3607(8) and 2.3571(7)Å. The ACN/ H2O solvent system (Table 3). Surprisingly, no side
54
two Ru–P bond lengths are similar and comparable with those products were obtained from over oxidation reaction.
55
found in other reported ruthenium complexes containing Based on the observations, next we focus our interest to
56
PPh3.[18] Since all the complexes display similar spectral proper- investigating the effects of the substitution in the complexes
57

ChemistrySelect 2018, 3, 3036 – 3041 3038  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3038/3041] 1
Full Papers

1 Table 3. Effect of substituent of catalyst.[a] Table 4. Oxidative Cleavage of Various Olefins with NaIO4 Catalyzed by
Complex 1[a]
2
3
4
Entry Complex Yield (%)
5
6 1 Complex 1 95
7 2 Complex 2 99
3 Complex 3 94
8
S.No Substrate Product Yield (%) TONb
9 [a] Conditions: 4-methylstyrene (0.2 mmol), catalyst (0.5 mol %), oxidant
(2.0 mmol) in solvents (5 mL, 1:1) for 1 h.
10
[b] Isolated yield. 1. 95 190
11
12 2. 96 192
13
using the above optimal conditions. We explain the catalyst
14 3. 99 198
activity of all the complexes 1–3 in the oxidation of olefins
15
(Table 3). In all the cases, excellent catalytic activity and
16 4. 99 198
selectivity towards the product were observed. All the com-
17
plexes displayed more or less similar catalytic activity, suggest-
18 5. 92 184
ing that there is no significant effect on the catalysis despite
19
the change of substituent on the ligand in the complexes. So
20
we have chosen complex 2 for oxidation of different alkenes 6. 90 180
21
and alkynes under above optimal conditions.
22
Subsequently, the catalytic oxidation of a range substituted
23 7. 92 184
terminal olefins are tested to investigate its scope and
24
limitations. The results were summarized in Table 4. All the
25 8. 97 194
products were isolated and purified with column chromatog-
26
raphy and confirmed by NMR data. Styrene and electron rich 4-
27
methylstyrene were oxidized to corresponding aldehydes with 9. 99 198
28
95 and 96% of yield (entries 1 and 2). Styrene containing
29
electron withdrawing group such as 4-chloro and 4- fluoro
30 10. 99 198
styrene afforded the corresponding aldehydes in excellent
31
yields under identical conditions (entries 3 and 4). But, 4-nitro
32
phenylethene underwent oxidation to give 4-nitrobenzalde- 11. 74 148
33
hyde with 92% yield (entry 5) due to delocalization of electron
34
in the nitro group (N=O bond). In addition, electron with-
35 12. 99 198
drawing group attached to styrene in the ortho position can
36
drop the yield to 90% (entry 6). The decrease of yield in the
37 13. 89 198
above entry may be due to steric effect. Based on these results,
38
substrates containing electron withdrawing groups (entries 1–
39
4) give the corresponding aldehydes with excellent yields than 14. 95 190
40
substrates with electron releasing groups. Moreover, this
41
methodology is more useful for the preparation of biologically
42 15. 98 195
active compound from oxidation of hetero aromatic olefin.
43
Accordingly, 2- vinyl pyridine afforded 92% to yield 2-pyridine
44 16. 95 190
carboxaldehyde without affecting any heterocyclic ring (en-
45
try 7). In addition, 2-vinyl naphthalene was employed under the 17. 99 198
46
same experimental conditions and afforded 97% yield of the
47 [a] Reaction conditions: Olefin (1 mmol), complex 2 (0.5 mol %), NaIO4
corresponding aldehyde (entry 8). The optimization is further
48 (2.0 mmol), CH3CN/H2O (1/1) (5 mL) at RT for 1 h.
extended to internal alkenes and other alkene derivatives to [b] TON = ratio of moles of product formed to moles of catalyst used.
49
check the efficiency of the present catalyst. Trans and cis
50
-stilbenes afforded 99% conversions to benzaldehyde (entries 9
51
and 10) respectively. Electron donating substitution on stilbene
52
make the olefin C=C become little steric nature and thereby substitution in a-position lowers the yield to 89% due to the
53
the cycloaddtion process will be disturbed and drop the yield steric effect on essential cycloaddtion process (entry 13). The
54
to 74% (entry 11). Methyl substituent on b-position of styrene protocol was further extended to internal alkynes. Interestingly,
55
afforded 99% of benzaldehyde (entries 12) without any un- the substrates phenyl acetylene and diphenyl acetylene are
56
desired dihydroxy or epoxide side products. But methyl efficiently catalyzed to their corresponding aldehydes with high
57

ChemistrySelect 2018, 3, 3036 – 3041 3039  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3039/3041] 1
Full Papers
yield (entries 14–15). Further, we explored the oxidative Electronic Supplementary Information (ESI) available: NMR
1
cleavage of cyclic aliphatic substrate in order to reveal the and CCDC reference numbers 1494359 and 1494360 contain
2
utility of the catalyst. Cyclohexene was readily cleaved to the supplementary crystallographic data for 1 and 2.
3
corresponding aldehyde with 99% yield (entry 16). Long chain
4
heptene gave the corresponding scission product in good yield
5 Acknowledgment
without any over oxidized acid side product (entry 17). The
6
unwanted side reactions like epoxidation, dihydroxylation or We sincerely thank the UGC-RFSMS, New Delhi, India. Also, we
7
allylic oxidations were not identified in this method. express our sincere thanks to DST-FIST, India for the instruments
8
Summing up, the catalytic action of the present system is facility at the School of Chemistry, Bharathidasan University,
9
appears to be better than similar type of catalysts reported Tiruchirappalli.
10
previously in terms of catalyst loading, conversions or yields
11
and reaction time. Ruthenium arene and ruthenium CNC pincer
12 Conflict of Interest
carbene complexes efficiently catalyzed the oxidation of olefins
13
under high catalyst loading in longer reaction time.[7,8] Shoair The authors declare no conflict of interest.
14
and co-workers have reported cis-[RuCl2(bipy)2].2H2O catalyzed
15
oxidative cleavage of alkenes to over oxidized acids in the Keywords: Environmentally benign oxidant · Open-Flask
16
presence of IO(OH)5 as the co-oxidant.[19] Dan Yang et al. have condition · Room temperature · Ru(II)thiosemicarbazone
17
developed a protocol to cleave olefins to carbonyl compounds complexes · Selective olefin oxidation
18
with high catalyst loading (3.5 mol %) of ruthenium trichlor-
19
ide.[20] Griffith and his co-workers have reported the oxidative [1] R. Stewart, R.Oxidation in Organic Chemistry; Wiberg, K., Ed.; Academic
20
cleavage of alkene and alkynes to their acid products using Press: New York, 1965; b) R. A. Sheldon, J. K. Kochi,Metal Catalyzed
21 Oxidations of Organic Compounds; Academic Press: New York, 1981;
RuCl3.nH2O-IO(OH)5 catalyst system.[21] Further, [cis-Ru(II)(dmp)2
22 c) M. Hudlicky, Oxidations in Organic Chemistry; American Chemical
(H2O)2]2 + have been proved to be effective catalysts in the
23 Society Monograph 186; American Chemical Society: Washington, DC,
catalytic oxidation of olefins to carbonyl products in presence 1990; d) R. C. Larock,Comprehensive Organic Transformations, 2nd ed.;
24
of NaIO4 as the oxidant using higher catalyst loading (3 mol % Wiley-VCH: New York, 1999, p1234; e) B. M. Trost, I. Fleming, Pergamon
25 Press: Oxford, 1991, Vol.7, p 541. H. Siegel, M. Eggersdorfer,Ullmann’s
).[22] It is to emphasize that our present catalyst works well at
26 Encyclopedia of Industrial Chemistry, 6th ed.; VCH, Weinheim, 2003,
room temperature by employing 0.5 mol% loading in 1 h.
27 Vol.18; f) F. Gasparrini, M. Giovannoli, D. Misiti, G. Natile, G. Palmieri, J.
Though we have not studied the mechanism of this reaction, Org. Chem. 1990, 55, 1323–1328; g) A. Baeyer, V. Villiger, Ber. Dtsch.
28
we believe that the mechanism of the reaction with ruthenium Chem. Ges. 1899, 32, 3625–3633; h) D. McReynolds, J. M. Dougherty, P. R.
29 Hanson, Chem. Rev. 2004, 104, 2239–2258; i) G. Zeni, R. Larock, Chem.
(II) complex involves the formation of ruthenium(VI)- dioxo
30 Rev. 2004, 104, 2285–2310.
species as reported by several workers.[7,23–25]
31 [2] a) K. Koike, G. Inoue, T. Fukuda, J. Chem. Eng. Jpn. 1999, 32, 295–299;
32 b) J. Andreusl, Chem. Revs. 1954, 54, 713.
[3] A. Sudalai, A. Khenkin, R. Neumann, Org. Biomol. Chem. 2015, 13, 4374–
33 Conclusion 4394.
34 [4] a) K. B. Sharpless, K. Akashi, J. Am. Chem. Soc. 1976, 98, 1986–1987;
New ruthenium(II)anthracenealdehyde thiosemicarbazone com-
35 b) B. R. Travis, R. S. Narayan, B. Borhan, J. Am. Chem. Soc. 2002, 124,
plexes have been synthesized and characterized. Single crystal 3824–3825; c) L. I. Simandi, T. L. Simandi, J. Mol. Catal. A: Chem. 1997,
36
X-ray diffraction studies of the complexes support the coordi- 117, 299–309; d) Y. H. Lin, I. D. Williams, P. Li, Appl. Catal. A: Gen. 1997,
37 150, 221–229; e) P. A. Ganeshpure, S. Satish, Tetrahedron Lett. 1988, 29,
nation of ligand is through hydrazinic nitrogen and thiolate
38 6629–6632; f) D. Xing, B. Guan, G. Cai, Z. Fang, L. Yang, Z. Shi, Org. lett.
sulfur and reveals the presence of distorted octahedral
39 2006, 8, No. 4, 693–696; g) A. Wang, H. Jiang, J. Org. Chem. 2010, 75,
geometry around Ru(II) ion. Further, the complexes show 2321–2326; h) C. Mi, X. G. Meng, X. H. Liao, X. Peng, RSC Adv. 2015, 5,
40
excellent catalytic performance in selective C = C bond scission 69487–69492; i) G. C. Angela, J. L. Xiao, J. Am. Chem. Soc. 2015, 137,
41 8206–8218; j) A. Dhakshinamoorthy, K. Pitchumani, Tetrahedron 2006, 62,
of a diverse range of alkenes to aldehydes and C  C bonds to
42 9911–9918; k) A. D. Chowdhury, R. Ray, G. K. Lahiri, Chem. Commun.
a-diketones in an aqueous/organic solvent system under mild
43 2012, 48, 5497–5499; l) F. Shi, M. K. Tse, M. M. Pohl, A. Bruckner, S. Zhang,
condition using NaIO4. The maximum isolated yields of the M. Beller, Angew. Chem. Int. Ed. 2007, 46, 8866–8868; m) M. Ma, Y. Yue,
44
product up to 99% are obtained using 0.5 mol% of catalyst W. Hua, Z. Gao, Appl. Catal. A. 2003, 251, 39–47.
45 [5] a) D. Yang, C. Zhang, J. Org. Chem. 2001, 66, 4814–4818; b) P. H. J.
loading. These studies significantly advance the scope of the
46 Carlsen, T. Katsuki, V. S. Martin, K. B. Sharpless, J. Org. Chem. 1981, 46,
oxidative cleavage of olefin reactions. The detailed mechanistic
47 3936–3938; b) S. Torii, T. Inokuchi, K. Kondo, J. Org. Chem. 1985, 50,
aspect of the oxidation reaction is under investigation. 4980–4982.
48
[6] J. K. Bera, P. Daw, R. Petakamsetty, A. Sarbajna, S. Laha, R. Ramapanicker,
49 J. Am. Chem. Soc. 2014, 136, 13987–13990.
50 Supporting Information Summary [7] a) E. A. Nyawade, H. B. Friedrich, B. Omondi, P. Mpungos, P. Organo-
51 metallics 2015, 34, 4922–4931; b) J. M. Gichumbi, H. B. Friedrich, B.
Electronic Supplementary Information (ESI) available: X-ray Omondi, J. Organomet. Chem. 2016, 808, 87–96.
52
structure, experimental data, and copies of 1H and 13C NMR [8] M. Poyatos, J. A. Mata, E. Falomir, R. H. Crabtree, E. Peris, Organometallics
53 2003, 22, 1110–1114.
spectra for all the olefin oxidation products are provided. This
54 [9] S. K. Gupta, K. Sahoo, S. Choudhury, J. Organometallics 2016, 35, 2462–
material is available free of charge via the Internet at http://
55 2466.
pubs.acs.org. [10] H. Okumoto, K. Dhtusuka, S. Banjoya, Synlett. 2007, 3201–3205.
56
[11] J. Chen, Chi-Ming Che, Angew. Chem. Int. Ed. 2004, 43, 4950–4954.
57

ChemistrySelect 2018, 3, 3036 – 3041 3040  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3040/3041] 1
Full Papers
[12] V. Kogan, M. M. Quintal, R. Neumann, Org. Lett. 2005, 7, 5039–5042. [23] a) E. A. Nyawade, H. B. Friedrich, B. Omondi, P. Mpungos, Organometallics
1 [13] a) D. X. West, S. B. Padhye, P. B. Sonawane, Struct. Bonding. 1991, 76, 1– 2015, 34, 4922–4931; b) J. M. Gichumbi, H. Friedrich, B. Omondi, J.
2 50; b) Fernandez, A. Fernandez, M. Lopez-Torres, H. Adams, J. Chem. Soc. Organomet. Chem. 2016, 808, 87–96; c) Y. Qing Zhong, H. Qiong Xiao,
3 Dalton Trans. 1999, 4193–4201. X. Yi Yi, Dalton Trans. 2016, 45, 1811; d) A. S. Goldstein, R. S. Drago, J.
[14] a) R. N. Prabhu, R. Ramesh, RSC Advances 2012, 2, 4515–4524; b) E. Chem. Soc. Chem. Commun. 1991, 21–22; e) P. Aguirre, R. Sariego, S. A.
4
Sindhuja, R. Ramesh, Organometallics 2014, 33, 4269–4278. Moya, J. Coord. Chem. 2001, 54, 401–413.
5 [15] a) Martinez, L. A. Adrio, J. M. Antelo, J. M. Ortigueira, M. T. Pereira, J. J. [24] a) P. Daw, R. Petakamsetty, A. Sarbajna, S. Laha, R. Ramapanicker, J. K.
6 Fernandez, A. Fernanadez, J. M. Vila, J. Organomet. Chem. 2006, 691, Bera, J. Am. Chem. Soc. 2014, 136, 13987–13990; b) J. M. Gichumbi, H. B.
7 2891–2901. Friedrich, B. Omondi, J. Organomet. Chem. 2016, 808, 87–96.
[16] N. Ahmed, J. J. Lewison, S. D. Robinson, M. F. Uttley, Inorg. Synth. 1974, [25] a) S. Goswami, A. R. Chakravarty, A. Chakravorty, Inorg. Chem. 1983, 22,
8
15, 48. 602–609; b) C. Ming Che, K. Y. Wong, W. Ho Leung, C. K. Poon, Inorg.
9 [17] J. C. Briggs, C. A. McAuliffe, G. Dyer, J. Chem. Soc. Dalton Trans. 1984, Chem. 1986, 25, 348–355; c) B. J. Hornstein, D. M. Dattelbaum, J. R.
10 423–427. Schoonover, T. J. Meyer, Inorg. Chem. 2007, 46, 8139–8145.
11 [18] J. G. Malecki, A. Maron, J. Kusz, Mendeleev Commun. 2015, 25, 103–105;
F. Basuli, S. M. Peng, S. Bhattacharya, Inorg. Chem. 2001, 40, 1126–1133.
12
[19] A. G. F. Shoair, R. H. Mohamed, Synth. Commun. 2006, 36, 59–64.
13 [20] D. Yang, Chi Zhang, J. Org. Chem. 2001, 66, 4814–4818. Submitted: January 18, 2018
14 [21] W. P. Griffith, E. Kwong, Synthetic communications 2003, 33, 2945. Revised: March 5, 2018
15 [22] V. Kogan, M. M. Quintal, R. Neumann, Org. Lett. 2005, 7, 5039–5042. Accepted: March 5, 2018
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

ChemistrySelect 2018, 3, 3036 – 3041 3041  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1811 / 108705
Freitag, 16.03.2018
[S. 3041/3041] 1

You might also like