You are on page 1of 15

Accepted Article

Title: Selective decomposition of cyclohexyl hydroperoxide using


homogeneous and heterogeneous Cr(VI) catalysts: Optimizing
the reaction by evaluating the reaction mechanism

Authors: Jessica Nadine Hamann, Marko Hermsen, Anna-Corina


Schmidt, Saskia Krieg, Jasmin Schießl, Dominic Riedel,
Joaquim Henrique Teles, Ansgar Schäfer, Peter Comba, A.
Stephen K. Hashmi, and Thomas Schaub

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: ChemCatChem 10.1002/cctc.201701909

Link to VoR: http://dx.doi.org/10.1002/cctc.201701909

A Journal of

www.chemcatchem.org
ChemCatChem 10.1002/cctc.201701909

FULL PAPER
Selective decomposition of cyclohexyl hydroperoxide using
homogeneous and heterogeneous Cr(VI) catalysts: Optimizing the
reaction by evaluating the reaction mechanism
Jessica Nadine Hamann,[a] Marko Hermsen,[a,b] Anna-Corina Schmidt,[a] Saskia Krieg,[c] Jasmin
Schießl,[d] Dominic Riedel,[e] Joaquim Henrique Teles,[e] Ansgar Schäfer,[b] Peter Comba,[c] A. Stephen
K. Hashmi,[a,d] Thomas Schaub*[a,e]

Accepted Manuscript
Abstract: In this study, known homogeneous and heterogeneous textile industry in the manufacture of yarn or as thin plastic film for
chromium(VI) catalyst systems were investigated with respect to the packaging purposes.[2] On an industrial scale, the polymer is
favored formation of cyclohexanone during the decomposition of produced via ring-opening polymerization of ε-caprolactam at
cyclohexyl hydroperoxide (CHHP). The focus was on mechanistic high temperatures.[1,2,4] The required starting material for the
studies using different spectroscopic methods as well as DFT synthesis of ε-caprolactam is cyclohexanone that reacts with
calculations to further optimize the reaction conditions. As in previous hydroxylamine in a condensation reaction to the corresponding
decomposition studies, a mechanism via the formation of a metal cyclohexanone oxime and through an acid-catalyzed Beckmann
alkylperoxido intermediate is probable. In situ spectroscopic studies rearrangement providing the desired ε­caprolactam (Scheme 1).[4]
revealed that in case of both the soluble and insoluble catalyst, the
selective decomposition happens via a non-radical, non-redox
mechanism at the Cr(VI) stage through the formation of a
cyclohexylperoxychromium(VI) complex. The proposed mechanism is
supported by thorough DFT calculations.

Introduction

Nylon 6 is one of the most widely used polymers and is produced Scheme 1. Industrial process for the production of Nylon 6 starting from
on a large industrial scale (~ 4 million t/a).[1-4] This important cyclohexanone.
polymer has a wide range of applications, for example in the
The well-established industrial process to obtain cyclohexanone
[a] Dr. J. N. Hamann, M. Hermsen, Dr. A.-C. Schmidt, Prof. Dr. A. S. K. is the autoxidation of cyclohexane, either non-catalyzed or
Hashmi, Dr. T. Schaub catalyzed by different transition metal compounds like soluble
Catalysis Research Laboratory (CaRLa) cobalt catalysts.[5-15] The liquid-phase oxidation of cyclohexane is
Im Neuenheimer Feld 584
conducted at temperatures up to 130 °C as well as high
69120 Heidelberg (Germany)
E-mail: thomas.schaub@basf.com pressures are required (125-165 °C, 8-15 bar). These conditions
[b] M. Hermsen, Dr. A. Schäfer are applied to shorten the retention time, in order to avoid
BASF SE, Quantum Chemistry & Molecular Simulation Catalysis oxidative side reactions. For the same reason, the cyclohexane
Carl-Bosch-Straße 38
conversion is limited to 10-12 % per cycle.[3] One common
67056 Ludwigshafen (Germany)
[c] S. Krieg, Prof. Dr. P. Comba intermediate in the oxidation procedure is cyclohexyl
Institute of Inorganic Chemistry & Interdisciplinary Center for hydroperoxide which further decomposes to a mixture of the
Scientific Computing corresponding ketone and alcohol.[16-18] In detail, in a radical
Heidelberg University
chain reaction cyclohexane is oxidized to cyclohexyl
Im Neuenheimer Feld 275
69120 Heidelberg (Germany) hydroperoxide by air at elevated temperature.[8,19] In the next step,
[d] J. Schießl, Prof. Dr. A. S. K. Hashmi intermediately formed cyclohexyl hydroperoxide decomposes to
Institute of Organic Chemistry an ~ 1:1 mixture of cyclohexanone (ketone K, CyO) and
Heidelberg University
cyclohexanol (A, CyOH), which is known as KA oil and is
Im Neuenheimer Feld 270
69120 Heidelberg (Germany) produced on a scale of more than 7 million tons per year (Scheme
[e] Dr. D. Riedel, Dr. J. H. Teles, Dr. T. Schaub 2).[8]
BASF SE, Synthesis and Homogeneous Catalysis
Carl-Bosch-Straße 38
67056 Ludwigshafen (Germany)
Supporting information for this article is given via a link at the end of
the document.

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
supported transition metal catalysts.[35,44-49] For example,
following the activity of soluble cobalt catalysts, poly(4-
vinylpyridine) (PVP) supported Co(II) catalysts have been
prepared and investigated for the oxidation of cyclohexane.[44-49]
In the same way, CrO3 immobilized on PVP was prepared for the
selective decomposition of cyclohexyl hydroperoxide and seemed
to be a promising candidate for this kind of reaction.[35] In 2003,
Scheme 2. Usual oxidation of cyclohexane to obtain KA oil and cyclohexyl Buijs et al. showed that homogeneous and heterogeneous Cr(VI)
hydroperoxide (CHHP). catalysts exhibit a comparable reactivity with respect to the
selective decomposition of cyclohexyl hydroperoxide towards
Cyclohexanol and cyclohexanone can be separated by distillation cyclohexanone.[41] The early work focused only on achieving high
and, depending on the requirements, the alcohol can be either conversions as well as high yields of cyclohexanone. However,
used for the production of adipic acid or dehydrogenated to so far there is no clear picture about the mechanism using Cr(VI)

Accepted Manuscript
cyclohexanone.[8] With the target of reducing the number of catalysts and furthermore, nothing is known about the oxidation
complexity in the production of cyclohexanone, it would be state of chromium at the end of the catalysis. To close this gap,
desirable to increase the selectivity towards the formation of the we focused on investigating the mechanism of decomposition of
ketone. Oxidation of cyclohexane without any metal catalyst cyclohexyl hydroperoxide. We investigated homogeneous and
forms cyclohexyl hydroperoxide as the major product.[8] Therefore, heterogeneous catalysts using different spectroscopic methods
a few patents and studies with regard on the development of and compared the experimental results with results from DFT
catalysts for the selective decomposition of cyclohexyl calculations. Especially, the structure of the reactive intermediate
hydroperoxide to cyclohexanone have been published.[20-34] is of interest. Based on the results, the dehydroperoxidation
During the last years, especially Cr, V, Co, Fe and Mn catalysts reaction can be understood and optimized for a possible
have been investigated for the selective dehydroperoxidation and application in industry. A possible mechanism is proposed based
seem to be the most promising.[27,28,30-33,35-37] In our previous work, on the combination of experimental and theoretical investigations.
we presented different vanadium(V) catalysts with dipicolinic acid
(dipic)-based ligands or Schiff base ligands for the selective
decomposition of different alkyl hydroperoxides.[21] Nevertheless, Results and Discussion
in the case of cyclohexyl hydroperoxide, a higher selectivity
towards the ketone could not be achieved. However, the obtained Catalyst screening
results point out that transition metal oxo species (M=O) seem to
be necessary in the design of an efficient catalyst. The terminal In order to find the most active chromium catalyst for the
oxo group supports a non-radical mechanism via a hydrogen dehydroperoxidation step and to investigate a possible influence
transfer from the secondary hydroperoxide to form the ketone in of the ligand framework, an extensive catalyst screening was
one step with a high overall selectivity.[21] Another mandatory performed under different reaction conditions. Therefore, several
aspect is the stability of the catalyst, even at temperatures higher known Cr(III) and Cr(VI) compounds were used in the
than 120 °C. As attractive systems, Cr(VI) compounds have been decomposition of neat cyclohexyl hydroperoxide (Scheme 3).
identified as promising leads, which fulfill these requirements and
enable the selective decomposition of cyclohexyl hydroperoxide
to cyclohexanone.[38-42] Especially chromium trioxide (CrO3) is
known to perform the desired reaction with very good
selectivity.[39,40] The use of neat CrO3 as catalyst to decompose
cyclohexyl hydroperoxide was first mentioned in a German patent Scheme 3. Selective decomposition of CHHP to cyclohexanone and water
in 1973.[39] using chromium catalysts.
In the last years, soluble CrO3 has been investigated in
comprehensive studies as a potential catalyst, but to the best of With focus on an industrial application, short reaction times at high
our knowledge, the Cr(VI) catalyst is not applied industrially in the temperatures are of particular interest and for this reason
production of cyclohexanone. The main drawbacks are the high temperatures up to 130 °C were included in the screening.
environmental risks associated with the use of Cr(VI), which is the
most toxic form of chromium.[41] An important aspect of designing
a catalyst is to cover all possibilities of increasing the efficiency
Table 1. Different chromium(III) and chromium(VI) compounds used as
and recyclability. Therefore, another approach to use the high catalysts in the decomposition of CHHP.
activity and in particular to benefit from high selectivity is the use
of heterogeneous chromium(VI) catalysts.[35,41,43] Besides Catalyst Molar Solv.a Condit. Conv. Select.
ratio [t, T] [%]b [%, K:A]b,c
chromium species supported on zeolites and mesoporous
Cr:CHHP
chromium aluminophosphates (CrAlPO-5),[30,41] the oxidation of
cyclohexane as well as the selective decomposition of cyclohexyl a) Cr(acac)3 1:184 Cy-d12 12 h, 42 90:10
hydroperoxide have been investigated extensively with polymer- rt

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
b) Cr(salen*) 1:181 CDCl3 2 h, 88 69:31 on (i) the used solvent and (ii) depending on the solubility of the
40 °C used catalyst at room temperature. In some cases an influence
on the selectivity was observed by increasing the temperature.
c) Cr2O3 1:183 CDCl3 72 h, 53 29:71
For example, using Cr(tBu-acac)3 as catalyst at 100 °C, a
60 °C
conversion of 85 % (after 20 min) was observed, but the
d) CrCl3 1:200 CDCl3 24 h, 91 55:45 selectivity decreased drastically to 81:9 % (see SI). Using
60 °C CrO3*py at elevated temperatures, a positive evolution in the
conversion could be observed along with small losses in
e) CrCl3*6H2O 1:287 CDCl3 2 h, 100 66:34
90 °C selectivity. Starting at 80 °C, a conversion of 88 % was observed.
Increasing the temperature by 10 °C, the conversion could be
f) Cr(tBu-acac)3 1:230 CDCl3 16 h, 80 96:4 increased to 94 %, whereas the same selectivity was obtained.
rt
Finally, the best combination of good selectivity and nearly full
g) Cr(acrylate)3 1:424 CDCl3 24 d, 13 30:70 conversion was detected by running the reaction at 100 °C.

Accepted Manuscript
rt Overall, reactions at higher temperatures are favored relating to
the industrial process, as the stream obtained directly from the
h) Cr(CF3-acac)3 1:298 CDCl3 8 d, 55 91:9
liquid-phase oxidation also has temperatures in the range of 100
rt
to 130°C. Therefore, it is beneficial from a process point of view if
i) Cr(Ph,Me- 1:1511 Cy-d12 12 h, 94 88:12 a selective dehydroperoxidation could be applied directly on this
acac)3 60 °C stream without changing the temperature (avoids additional heat
exchangers).
j) K2Cr2O7 1:766 CDCl3 6 h, 44 62:38
60 °C As shown in Table 1, the best results regarding high
conversion and high selectivity towards the ketone as well as
k) CrO3*bipy 1:415 CDCl3 30 min, 64 56:44 short reaction times at elevated temperatures were obtained
120 °C
using a soluble CrO3-pyridine complex as catalyst. Despite the
l) CrO3*phen 1:388 CDCl3 30 min, 45 58:42 excellent results observed with pure CrO3 in CDCl3 at room
120 °C temperature (Table 1, entry m), the use of a chromium-pyridine
complex as catalyst was necessary because neat CrO3 is
m) CrO3 1:400 CDCl3 10 min, 100 98:2
insoluble in the required non-polar solvents and addition of CrO3
rt
to cyclohexyl hydroperoxide at room temperature leads to a highly
n) CrO3*py 1:441 Cy-d12 2 h, 88 96:4 exothermic reaction. At higher temperatures, the precipitation of
80 °C an insoluble black solid was observed, due to the rapid formation
of unknown inactive chromium species. The results lead to the
o) CrO3*py 1:477 Cy-d12 2 h, 94 96:4
90 °C assumption that (i) a Cr(VI) center is necessary but (ii) a fancy
ligand framework is not required for an efficient chromium catalyst.
p) CrO3*py 1:512 Cy-d12 1 h, 99 95:5 Moreover, using bidentate N-donor ligands like 2,2'-bipyridine
100 °C
(bipy) or 1,10-phenanthroline (phen), the conversion and
q) CrO3*py 1:380 Cy-d12 45 min, 99 95:5 selectivity decreased drastically. The dehydroperoxidation with
120 °C Cr(VI) works equally at room and elevated temperatures. Figure
1 shows as an example the 1H NMR spectra measured before
r) CrO3*py 1:560 Cy-d12 30 min, 98 90:10
and after the catalytic run using CrO3*py at 100 °C. After one hour,
130 °C
the substrate is completely decomposed and only small amounts
s) CrO3*py 1:398 Cy-d12 30 min, 98 84:16 of cyclohexanol are detected.
150 °C In contrast to our previous studies with vanadium(V)
complexes, the yield of cyclohexanone is not significantly
[a] 0.5 mL deuterated solvent. influenced by higher reaction temperatures. The slightly lower
[b] Results based on NMR data with naphthalene as internal standard. selectivity observed when the reaction is run at 130 °C is probably
[c] K = cyclohexanone, A = cyclohexanol. due to the competing uncatalyzed thermal decomposition of
* A modified salen ligand was used. CHHP via a radical mechanism. This non-catalyzed unselective
decomposition lowers the yield of the desired cyclohexanone. In
a separate experiment, the half-life for thermal decomposition of
The dehydroperoxidation reactions of CHHP were performed in CHHP at 130 °C was determined as 29 min.
different deuterated solvents. In a general procedure, a solution
containing the catalyst, substrate and naphthalene as internal
standard was prepared in a NMR tube. Depending on the reaction
conditions, the tube was heated or stored at room temperature. In
case of strong exothermic reactions, the NMR tube was stored in
an ice bath. The reaction mixture was analyzed by NMR
spectroscopy. The reaction temperatures were chosen depending

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
At 5 °C a first UV/vis spectrum was measured, which is shown
After 1 h@100 °C CyO in Figure 2 (black line). As expected for Cr(VI), no intensive
absorption band is observed. After addition of stoichiometric
amounts of CHHP, a UV/vis spectrum was recorded every
10 seconds. During the first two minutes of reaction, the rapid
* formation of an absorption band at 480 nm was observed.
CyOH According to the literature, this band is assigned to a peroxy to
Before
chromium charge transfer transition and indicates the formation
of the expected cyclohexylperoxychromium(VI) complex.[50,51]
After the rapid formation of the peroxychromium(VI) complex, the
CHHP
formed intermediate decomposed at a slightly slower rate. This
* behavior was observable by a shift of the absorption band to lower
wavelength (420 nm) within 10 min (Figure 2, inset black line). An

Accepted Manuscript
additional measurement after 48 hours revealed that the complex
had been completely decomposed (inset red line) to another
chromium species with different intensity of absorption. Boitsov et
Figure 1. 1H NMR spectra before and after the catalytic run using 1 eq. CrO3*py
al. also observed the same behavior in case of the formation of a
as catalyst at 100 °C for 1 h, measured in cyclohexane-d12. Signals marked with tert-butylperoxychromium complex.[51]
an asterisk belong to the internal standard (naphthalene).

Mechanistic studies

Using the optimized catalyst (CrO3-pyridine complex),


different spectroscopies were used to gain insight into the
chromium-catalyzed dehydroperoxidation. Initially, in situ UV/vis
2 min
measurements were performed to identify the oxidation state of
the metal center during catalysis. A solution of the catalyst in
10 min
cyclohexane was prepared and the UV/vis spectrum was
48 h
measured at ambient temperatures. The absorption spectrum of
the pure catalyst shows no intensive absorption bands in the
range between 300 and 1000 nm, which is in agreement with the
oxidation state +6 of chromium (see SI). After the addition of
500 eq. of substrate at room temperature, the fast catalytic
reaction was followed for 60 min. The intensity of the weak
absorption at ~ 450 nm decreases due to the coordination of the
substrate to the metal center, but during the following catalytic
decomposition of the cyclohexyl hydroperoxide, no further
absorption bands were observed. Especially, in the interesting
area around 600 nm no new absorption bands appear. This
implies that the oxidation state of chromium does not change
Figure 2. UV/vis spectra detected during the rapid stoichiometric reaction of a
during the catalytic decomposition reaction. At the end of the 0.93 m M solution of CHHP in α,α,α-trifluorotoluene and 1 eq. CrO3*py at 5 ° C.
reaction, the oxidation state is still +6, as in case of a reduction to Inset: Spectrum of formed cyclohexylperoxychromium(VI) complex after 2 min
Cr(III) two spin allowed d-d-transitions of the d3 configuration in (blue line); decomposition and shift to lower wavelength after 10 min (black line)
the range of 420 - 450 nm and 580 - 620 nm are expected. and full decomposition after 48 h (red line); l = 1 cm.
Based on this first indication concerning the oxidation state of
the chromium catalyst, the reactive intermediate was The cyclohexylperoxychromium(VI) complex formed when a
characterized by low-temperature measurements. In accordance stoichiometric amount of CHHP is added decomposes slowly to
with our proposed mechanism for the vanadium(V) catalysts,[21] the ketone and CrO2(OH)2 over time. In the presence of water, the
the formation of an alkylperoxychromium(VI) complex was formation of chromic acid is favored and stabilized in the aqueous
expected. Following the procedure by Boitsov et al. to detect a environment, especially at low temperatures. As Boitsov et al.
tert-butylperoxychromium(VI) complex, in situ UV/vis showed in their study, an acidic environment influences the
measurements were performed at low temperatures.[51] Therefore, stability of the investigated tert-butylperoxychromium(VI) complex
a solution of CrO3*py in anhydrous α,α,α-trifluorotoluene was drastically.[51] This implies that the formation and stability of the
prepared and cooled to 5 °C. The non-coordinating solvent cyclohexylperoxychromium(VI) complex is influenced by the
α,α,α-trifluorotoluene (BTF) was chosen to further slowdown the strongly acidic reaction mixture. The following experiment
rate of reaction.[51] supported this assumption: 10 equivalents of cyclohexyl
hydroperoxide were added to the solution containing the

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
decomposed cyclohexylperoxychromium(VI) complex at 5 °C, but cyclohexylperoxychromium(VI) complex and is catalytically
the formation of a new peroxy complex could not be observed. inactive. Addition of an excess of cyclohexyl hydroperoxide did
The decomposition of the substrate to cyclohexanone was not not form the cyclohexylperoxychromium(VI) complex and
observed with the inhibited catalyst at low temperatures. On the decomposition of the cyclohexyl hydroperoxide was not observed.
other hand, adding 100 eq. cyclohexyl hydroperoxide to the The formation of the reactive intermediate
inhibited cyclohexyl-peroxychromium(VI) complex and running a (cyclohexylperoxychromium(VI) complex) was also detected by
catalytic reaction at 120 °C, the selective formation of NMR spectroscopy. Due to the observed rapid formation and
cyclohexanone was observed. That implies, that the regeneration decomposition of the reactive intermediate, measurements were
of the cyclohexylperoxychromium(VI) complex is suppressed at performed at -20 °C in deuterated toluene (toluene-d8). To gain
low temperatures but takes place at higher temperatures. By insights into the coordination of cyclohexyl hydroperoxide to the
using an excess of CHHP like under the conditions of the catalytic chromium center, in the first step a 1H NMR spectrum of pure
CHHP decomposition, the equilibrium seems to be shifted to the CHHP was measured at -20 °C (Figure 3 lower part). The focus
active form of the catalyst. was put on the proton signal at 3.72 ppm (red label), which was

Accepted Manuscript
Another experiment was also performed by adding assigned to the CH-group next to the OOH group. Immediately
2.5 equivalents of substrate to 1 equivalent of Cr(VI). The same after addition of stoichiometric amounts of the catalyst, a
absorption band as described in Figure 2 (blue line) was observed, downfield shift to 4.09 ppm of this signal was observed (Figure 3
but its formation and subsequent decomposition happened twice upper part).
as fast. In further experiments, the stability of the reactive
cyclohexylperoxychromium(VI) complex was investigated by the
addition of water. To the reactive intermediate one equivalent of
water was added that leads immediately to the formation of
undefined chromium(VI) species. NMR investigations of the
reaction mixture showed, that cyclohexanone was not formed
Addition of CrO3*py @ -20°C
during the water-induced rapid conversion at low temperatures.
OOCH
This implies that the catalyst is inhibited by the presence of water
at low temperatures. To further investigate the influence of formed OH
water during the catalytic decomposition of CHHP, a catalytic
reaction was performed in the presence of a 25-fold excess of
water (7500 eq according to Cr(VI)). The obtained results showed
OOH
that larger quantities of water are inhibiting the catalytic activity of
Pure CHHP @ -20°C
Cr(VI) indicated by the observed lower conversion (99 % vs.
46 %) and selectivity (95 % vs. 71 % ketone, for further details OOCH
see SI). This aspect is in line with the reported removal of
formed water from the reaction to avoid catalyst inhibition.[41]
The addition of two equivalents cyclohexanol to the
stoichiometrically formed cyclohexylperoxychromium(VI)
(480 nm) complex led to a shift of the charge transfer transition to
lower wavelength (455 nm) but no decrease in intensity. The blue Figure 3. Low temperature 1H NMR spectra of pure CHHP (lower part) and after
shift may be due to ligand exchange (CHHP to cyclohexanol).[51] the addition of stoichiometric amounts of CrO3*py (upper part), measured in
This formed cyclohexyl chromate complex is more stable than the toluene-d8. Marked are significant changes and shifts due to the formation of a
cyclohexylperoxychromium(VI) complex.
Table 2. Comparison of experimental and calculated 1H NMR shifts,
which are observable during the formation of the
cyclohexylperoxychromium complex. This is induced by coordination of the peroxide group to the
metal center, leading to a reduction of the electron density and
NMR shifts / ppm Experimental Calculated[51] the proton becomes more deshielded. The broadening of the
signal, due to the metal environment, supports this aspect.
Pure CHHP Furthermore, the 1H signal of the OOH group at 5.83 ppm
disappeared, and a new broad signal at 11.64 ppm appeared
OOH 5.83 6.62
(blue label). In order to verify the observed changes in the NMR
OOCH 3.72 3.84 spectra, DFT calculations of the expected shifts were carried
out.[52] A comparison of the relevant experimental and calculated
Coordinated 1
H NMR shifts is provided in Table 2.
species
CrO2(OH)(OOCy) Although there is a systematic shift between measured and
calculated shifts, there is an excellent linear correlation between
OH 11.64 11.82 measured and calculated shift, with a slope of 1 and a correlation
coefficient r2= 0.992. The observed downfield shift of relevant
OOCH 4.09 4.19
signals is induced by the formation of the

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
cyclohexylperoxychromium(VI) intermediate. In agreement with cyclohexyl hydroperoxide was investigated with a focus on the
the calculated shifts, the OOCH-signal shifts by 0.37 ppm. Based mechanism as well as leaching properties.
on the calculated results, the new broad signal at 11.64 ppm is In accordance to Lončarević et al., insoluble poly(4-vinylpyridine)
assigned to the OH group on Cr(VI). Finally, the disappearance of supported chromium(VI) catalysts were prepared by adding CrO3
the hydroperoxide group signal further supports that the substrate to a suspension of PVP in water.[35] To investigate the influence
is coordinated to the hexavalent chromium center. In the 13C of the metal loading, catalysts with chromium(VI) contents in the
spectra, the calculation predicts only small shifts in the range of range of 0.9 until 7.3 wt. % were used. Using the same chromium
0.1-0.7 ppm upon coordination. Still these are in good agreement to substrate molar ratio (1:470) in the following catalytic studies at
with the experimentally obtained 13C NMR data (for details see SI). 60 °C, the influence of the chromium loading was determined. As
Besides UV/vis and NMR measurements, the dehydro- shown in Figure 4, the specific activity of the chromium catalyst
peroxidation reaction was also investigated using in situ IR decreases with an increasing metal content on PVP.
spectroscopy. At different temperatures, the reaction of
cyclohexyl hydroperoxide and CrO3 was monitored in order to

Accepted Manuscript
detect possible vibrations of the cyclohexylperoxychromium(VI)
intermediate. Therefore, the reaction was performed in both a
stoichiometric as well as in a catalytic mode. Relevant but weak
chromium(VI)-oxygen vibrations should be observable in the
range of 700 cm-1 to 900 cm-1, but despite comprehensive studies,
no characteristic vibrations, which belong to the reactive
intermediate, could be detected.[50] However, during the catalytic
run, an increasing formation of cyclohexanone could be observed
due to the decomposition of CHHP (see SI). After the addition of
1 eq. CrO3 to 250 eq. cyclohexyl hydroperoxide at -74 °C and
warming to room temperature, an intensive vibration appears in
the range of 1700 to 1750 cm-1, which increases in intensity with
reaction time. Due to the characteristic range of carbonyl Figure 4. Comparison of the observed CHHP conversion after 80 min at 60 °C
stretches, this new intensive band is assigned to the C=O stretch at the same Cr:CHHP ratio of 1:470 in cyclohexane-d12 using catalysts with
of emerging cyclohexanone (see SI). different Cr(VI) loadings on PVP.
With respect to the obtained mechanistic insights, it is
possible to propose that the dehydroperoxidation of cyclohexyl For example, using CrPVP with a chromium content of 7.3 wt. %,
hydroperoxide happens via (i) the formation of an the conversion of CHHP was only 45 % after 80 min at 60 °C (see
alkylperoxychromium(VI) intermediate and (ii) Cr(VI) is not SI). A conversion of 98 % was achieved only after 5.5 h, leading
reduced during the catalysis. to a ketone:alcohol:dicyclohexyl peroxide ratio of 55:32:13 (Table
3). Using the lowest loaded CrPVP catalyst, a nearly full
Heterogeneous Cr(VI) catalysis conversion was obtained after 2.5 h with a selectivity of 74 %
ketone. With a higher chromium content, the rate of reaction
In the context of a potential industrial application, a dissolved decreases and the formation of dicyclohexyl peroxide could be
Cr(VI) catalyst would require an elaborated recycle concept and detected by NMR spectroscopy. Due to a lower catalytic activity,
efficient disposal and treatment to remove or reduce the toxic uncatalyzed radical reactions are favored to form cyclohexyloxy
Cr(VI) species in the work-up, implicated its toxicity. In order to radicals, which induce the formation of this coupled side-product
use the desirable properties of chromium(VI) catalysts, a possible (Figure 5).
approach is the use of heterogeneous chromium(VI) catalysts.[35]
Especially the use of Cr(VI) immobilized on the well-known poly(4-
vinylpyridine) (PVP) presents the simplest approach to
heterogenize the CrO3*py complex used above.[53-56] Poly(4-
vinylpyridine) supported Cr(VI) complexes have already been Figure 5. Structure of the side-product dicyclohexyl peroxide, whose formation
studied as catalysts for the decomposition of cyclohexyl increases from 6 % to 14 % (after 80 min )with a Cr(VI) loading of 7.3 wt. %.
hydroperoxide. In 2010, Lončarević et al. presented a study using
poly(4-vinylpyridine)-supported cobalt(II) and chromium(VI) Table 3. Catalytic decomposition reaction of CHHP using CrPVP with different
catalysts in the oxidation of cyclohexane and decomposition of Cr loadings.
cyclohexyl hydroperoxide under mild conditions.[35] In the case of
the dehydroperoxidation, CrPVP showed a higher selectivity Catalyst Molar Solv.a Condit. Conv Select.
ratio [t, T] . [%,
towards cyclohexanone, whereas the activity of CoPVP was CrPVP Cr:CHHP [%]b K:A:CyOOCy]b
higher. Possible effects due to leaching of metal ions were only ,c

investigated for the cobalt catalysts. Regarding the mechanism


using PVP supported catalysts, no clear picture was given. To 0.9 wt.% 1:469 Cy-d12 2.5 h, 98 74:20:6
60 °C
address this point, the CrPVP catalyzed decomposition of

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
2.6 wt.% 1:471 Cy-d12 3 h, 98 75:19:6 CHHP was performed at different temperatures and the reaction
60 °C mixtures analyzed by NMR spectroscopy using naphthalene as
internal standard. These results are listed in Table 4. Reaction
3.5 wt.% 1:470 Cy-d12 3.5 h, 98 65:26:9
temperatures between 80 °C and 150 °C were chosen which are
60 °C
in the range of temperatures used in the dehydroperoxidation in
7.3 wt.% 1:469 Cy-d12 5.5 h 98 55:32:13 industrial plants.[8] With increasing temperature, the rate of
60 °C reaction also increases. At 150 °C the reaction time required for
full conversion could be reduced to 30 min.
[a] 0.5 mL deuterated solvent.
[b] Results based on NMR data with naphthalene as internal standard.
Table 4. Catalytic decomposition reaction of CHHP using CrPVP (2.6 wt% Cr-
[c] K = cyclohexanone, A = cyclohexanol, CyOOCy = dicyclohexyl peroxide. loading) at different temperatures.

An explanation of these results is the formation of less- Catalyst Molar Solv.a Condit. Conv. Select.
reactive dichromate species due to an overloading with chromium. [%]b [%, K:A]b,c

Accepted Manuscript
ratio [t, T]
In order to clarify this point, a dichromate-loaded quaternized PVP Cr:CHHP
was prepared. The quaternized PVP was used to rule out
t) CrPVP 1:545 Cy-d12 2 h, 99 98:2
coordination of the chromium in the dichromate to the N-donor 80 °C
ligand. The quaternized PVP precursor was prepared by
methylating PVP using a well-known procedure.[57] The u) CrPVP 1:412 Cy-d12 2 h, 99 98:2
90 °C
successful quaternization was followed by IR spectroscopy to
detect the disappearance of the vibration of the pyridine ring at v) CrPVP 1:647 Cy-d12 1 h, 99 98:2
~ 1600 cm-1 and the appearance of a new one at ~ 1640 cm-1, 100 °C
which belongs to the quaternized pyridine ring.[58] The methylated
w) CrPVP 1:589 Cy-d12 45 min, 98 92:8
PVP was then used to prepare a PVP-supported dichromate
120 °C
catalyst (CrMePVP, Scheme 4) containing 3.2 wt. % Cr.[59]
x) CrPVP 1:611 Cy-d12 30 min, 98 91:9
130 °C

y) CrPVP 1:450 Cy-d12 30 min, 98 82:18


150 °C

[a] 0.5 mL deuterated solvent.


[b] Results based on NMR data with naphthalene as internal standard.
[c] K = cyclohexanone, A = cyclohexanol.

Scheme 4. Reaction overview to obtain the dichromate catalyst supported on However, at temperatures higher than 130 °C, the selectivity
MePVP. decreases due to the competing thermal decomposition of CHHP.
The best results were obtained using moderate reaction times at
When running a similar catalytic decomposition reaction using 100 °C. The same behavior was observed running the reaction
CrMePVP as catalyst, comparable results were obtained to the with the homogeneous CrO3*py catalyst. A comparison of the
experiment with the highest loading of CrO3: 55 % conversion of evolution of selectivity depending on the reaction temperature is
CHHP and an increased formation of cyclohexyl peroxide (see SI). given in Table 5.
In the literature, different dichromate PVP species are known as
oxidizing agents, but only for the oxidation of alcohols. However,
in the case of cyclohexanol, long reaction times (more than 20 h)
and high temperatures are required for its oxidation yielding 72 %
cyclohexanone.[59] Therefore, it can be assumed that the
decreasing activity at a higher chromium content is connected
with the formation of dichromate species that are less effective for
the catalytic decomposition of cyclohexyl hydroperoxide.
To gain insights into the CrPVP-catalyzed dehydro-
peroxidation, different studies including kinetic studies were
performed. Measurements with the 0.9 wt. % and 2.6 wt. %
catalysts at 100 °C revealed a slightly higher conversion (after
1 h) by using the higher loaded CrPVP (99 % with 2.6 wt.% vs.
97 % with 0.9 wt.%). The selectivity is the same for both systems
(98:2) Therefore, in the following experiments, the catalyst used
had a chromium loading of 2.6 wt. %. The decomposition of

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
Table 5. Comparison of the evolution of selectivity depending on the Table 6. Reusability of CrPVP in the catalytic decomposition reaction of
temperature using homogeneous and heterogeneous Cr(VI) catalysts. CHHP over 10 cycles at 100 °C.

CrO3*py Conv. Select. CrPVP Conv. Select. Run Catalyst Molar Solv.a Condit. Conv. Select.
[%]b [%, K:A]b,c [%]b [%, K:A]b,c ratio [t, T] [%]b [%, K:A]b,c
Cr:CHH
80 °C 88 96:4 80 °C 99 98:2 P

90 °C 94 96:4 90 °C 99 98:2 1. CrPVP 1:507 Cy-d12 1h 99 98:2


100 °C
100 °C 99 95:5 100 °C 99 98:2
2. CrPVP 1:531 Cy-d12 1h 98 97:3
120 °C 99 95:5 120 °C 98 92:8 100 °C

130 °C 98 90:10 130 °C 98 91:9 3. CrPVP 1:498 Cy-d12 1h 98 95:5


100 °C

Accepted Manuscript
150 °C 98 84:16 150 °C 98 82:18
4. CrPVP 1:487 Cy-d12 1h 95 93:7
100 °C
In previous studies on heterogeneous chromium(VI) catalysts,
leaching of active chromium ions was the reason for the 5. CrPVP 1:501 Cy-d12 1h 95 91:9
100 °C
observable selective catalysis.[41] Due to the high toxicity of Cr(VI),
special interest was put on a possible leaching behavior of CrPVP. 6. CrPVP 1:505 Cy-d12 1h 93 91:9
In a first attempt to rule out leaching of chromium ions, the 100 °C
colorless product solution was investigated using ICP-MS. The
7. CrPVP 1:517 Cy-d12 1h 92 90:9
chromium concentration was below the detection limit of the
100 °C
method used (<0.1 mg kg-1). Further evidence to exclude
leaching of active chromium was obtained by running a split test. 8. CrPVP 1:499 Cy-d12 1h 90 89:11
Therefore, a dehydroperoxidation reaction was performed at 100 °C
55 °C to slow down the decomposition of CHHP. After 25 min, the
9. CrPVP 1:506 Cy-d12 1h 87 87:13
hot clear supernatant reaction mixture was divided into two parts. 100 °C
Both samples, a) with catalyst and b) without catalyst were heated
to 55 °C for further 40 min. The following NMR spectroscopic 10. CrPVP 1:503 Cy-d12 1h 85 85:15
100 °C
investigation revealed that in case of sample a) the substrate was
further decomposed to the ketone (for details see SI). In contrast,
[a] 0.5 mL deuterated solvent.
in sample b) the decomposition of CHHP stagnated during further
heating to 55 °C. Moreover, no changes in the compared NMR [b] Results based on NMR data with naphthalene as internal standard.

spectra are observable (see SI). The same procedure was [c] K = cyclohexanone, A = cyclohexanol.
performed to exclude possible leaching behavior at elevated
temperatures by running a split test at 100 °C. To slow down the
By reusing the catalyst ten times, the selectivity towards
fast reaction, a Cr:CHHP ratio of 1:1100 was applied. After 2 min
cyclohexanone decreased by 13 %, whereas the conversion
of reaction, the hot supernatant solution was separated and
declined by 14 %. Following the recycle tests using NMR
investigated as described above. In case of sample a) the
spectroscopy, the formation of small traces of dicyclohexyl
decomposition further happened. In case of sample b) the
peroxide started after five cycles (see SI).
selective decomposition mainly stagnated, but due to the
uncatalyzed self-decomposition of CHHP at elevated
Kinetic studies using CrO3*py and CrPVP catalysts
temperatures, traces of dicyclohexyl peroxide were detected. The
obtained results underline that active chromium ions were not
The results with homogeneous and heterogeneous
leaching during the catalysis, even at elevated temperatures, and
chromium(VI) catalysts show that both types of catalysts
as expected, the decomposition happens by a purely
decompose cyclohexyl hydroperoxide selectively to
heterogeneous pathway.
cyclohexanone. The mechanistic studies concerning the
For an industrial application, the recyclability of the CrPVP
homogeneous pathway strongly suggest dehydroperoxidation
catalyst is also a matter of interest and a reusability test was
occurs via the formation of a cyclohexylperoxychromium(VI)
conducted in ten cycles with the CrPVP at 100 °C (each run 1 h).
species.
After running a catalytic decomposition of CHHP, the supernatant
To compare both catalytic systems and to gain insights into
clear solution was removed and the catalyst was washed with
the reaction order, kinetic studies using 1H NMR spectroscopy
cyclohexane (3 x). Fresh substrate was added and a further
were performed. The reaction was followed by 1H NMR in
catalytic dehydroperoxidation was performed. As shown in Table
cyclohexane-d12 at temperatures between 50 °C and 66 °C to
6, the reusability could be confirmed with slight loss in activity (see
obtain the rate constants k at different temperatures. Due to the
Table 6).
slower reaction rate, the catalytic decomposition of cyclohexyl

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
hydroperoxide could conveniently be monitored during the first
80 minutes of reaction. The 1H NMR spectra using CrPVP as
catalyst at 50 °C are shown in Figure 6.

80 min
*

60 min *

40 min *

Accepted Manuscript
20 min *
Figure 7. Plot of ln[A]t/[A]0 vs. time for the decreasing concentration of CHHP
during the catalytic conversion using CrPVP as catalyst. The rate constants k
5 min are obtained from the slope.
*

After 80 min of reaction, the maximum conversion was


determined to 75 % at 339 K. Due to the incomplete
decomposition of CHHP, the relative small amounts of formed
Figure 6. 1H NMR spectra recorded following the catalytic dehydroperoxidation
water are negligible and have no influence on the reactions rates
of CHHP using CrPVP at 50 °C for 80 min, measured in cyclohexane-d12.
Signals marked with an asterisk belong to the internal standard (naphthalene). regarding possible inhibition effects. As shown above, a 25-fold
As shown in Figure 6, with increasing conversion of CHHP, excess of water is necessary to see a significant inhibition effect
the concentration of cyclohexanone increases. The reaction rates under the conditions of the catalysis. According to transition state
for the decomposition of cyclohexyl hydroperoxide at different theory, the activation parameters ΔH‡ and ΔS‡ were determined
temperatures were obtained using the differential equation from the temperature dependence of the obtained rate constants
describing first-order kinetics.[60] Both catalytic systems k. A representative Eyring plot, ln (k/T) vs. 1/T for the CrPVP
(homogeneous and heterogeneous) follow first-order kinetics as system, is given in Figure 8.
revealed by the linear plots in Figure 7 and the determined rate
constants for the decomposition of CHHP are given (Table 7).

Table 7. Rate constants for the decomposition of cyclohexyl


hydroperoxide into cyclohexanone following first-order-kinetics.

CrO3*py T/K k / 10-5 s-1 CrPVP T/K k / 10-5 s-1

323 15.6 ± 1.53 323 16.1 ± 1.01

331 24.2 ± 1.49 331 27.3 ± 1.65

339 29.6 ± 2.05 339 47.2 ± 1.88

The corresponding linear plots of ln {[CHHP]/[CHHP]0 versus


time, at 323, 331 and 339 K for the CrPVP system are shown in
Figure 7. The values of the rate constants k were obtained from Figure 8. Eyring plot for the decomposition of CHHP to cyclohexanone using
the slope of the fitted linear equation. A comparable plot for the CrPVP as catalyst.
soluble CrO3*py system is given in the SI. In order to confirm that
the decomposition follows a first-order kinetics depending on the Using the Eyring plot, the value of ΔH‡ can be obtained from
catalyst concentration, further studies with different chromium the slope whereas the value of ΔS‡ is obtained from the intercept.
concentrations were performed at 323 K. The results based on For the decomposition reaction using CrPVP, the activation
the linear plots confirm a clear correlation between the reaction parameters were estimated to ΔH‡ = 59.2 ± 2 kJ mol-1 and
rates k and the chromium concentration. This aspect leads to the ΔS‡ = -137.3 ± 5 J K-1mol-1. At 323 K, the Gibbs free energy of
conclusion that the dehydroperoxidation follows first-order activation was calculated to be ΔG‡ = 102.8 ± 3 kJ mol-1.
kinetics depending on (i) substrate and (ii) chromium Interestingly, for the homogeneous CrO3*py system (see SI for
concentration (see SI). details), the Gibbs free energy of activation was nearly identical
(ΔG‡ = 102.7 ± 8 kJ mol-1) although the activation enthalpy and
entropy are different (ΔH‡ = 34.8 ± 7 kJ mol-1 and ΔS‡ = -213.7 ±

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
22 J K-1mol-1). The strongly negative entropy of activation
additionally shows that the transition state has a higher degree of ∆G
organization than the reactants, which is also in accordance with kJ/mol O O O
Cr H
HO O
the proposed mechanism. O
H

The values for the activation parameters indicate that the TS4 O H
O O O RO
Cr
homogeneous as well as the heterogeneous Cr(VI) systems HO Cr
H O O O
O O O TS6
follow the same mechanism for the catalytic decomposition of H
H H

cyclohexyl hydroperoxide to cyclohexanone, and both reactions TS3


can be described by a kinetic which is first-order in both CHHP OOH

and Cr. To gain further insight into the mechanism as well as the H2O
R= H 83.7
activation parameters, comprehensive DFT studies were O O O
H 83.7
OCy 85.3 (TS7)
HO Cr
performed. O O
H
O
OOCy 87.8 (TS8)
H

Accepted Manuscript
O
TS2
Computational investigations HO
Cr
OH
O
O
1 Cr O
DFT calculations were used to further elucidate the mechanism of HO
O
O

the chromium-catalyzed dehydroperoxidation. Based on the 5


obtained computational results, the following mechanism can be O O

proposed for the dehydroperoxidation reaction using HO


Cr
O
OH +

chromium(VI) catalysts. 1
The key step is a hydrogen transfer from the peroxide
α-carbon to the oxo ligand on the chromium(VI) center with
Figure 9. Mechanism for the dehydroperoxidation of cyclohexyl hydroperoxide
simultaneous cleavage of the oxygen-oxygen bond. This by chromic acid and derivatives, energies given are free enthalpies in kJ mol-1.
mechanism was suggested earlier by Buijs et al. who studied this The rate-determining step is the release of ketone from the alkyl peroxide
complex 5. Addition of cyclohexyl hydroperoxide to the free acid is greatly
particular step in the reaction at the B3LYP/6-31G** level.[41] Our accelerated by proton relay agents such as cyclohexanol or water.
investigation shows that hydrogen transfer and O-O bond
cleavage is indeed the rate-determining step, considering The barriers for the cyclohexyl ester or dialkyl peroxide are
different variants of chromic acid esters and alkyl peroxide very similar (85.3 kJ mol-1 and 87.8 kJ mol-1, respectively). In the
species that may be present in the reaction mixture as well as presence of pyridine, a chromium(VI) complex with distorted
complexes and adducts that can be formed in the presence of trigonal planar coordination geometry can be formed, resembling
pyridine (as a model for poly(4-vinylpyridine), PVP). As given in the Collins reagent’s structure.[61] The activation barriers of the
Figure 9, starting from chromic acid 1, addition of cyclohexyl dehydroperoxidation for the pentacoordinated complexes with
hydroperoxide can proceed through a proton relay mechanism pyridine depend stronger on the second ligand (due to interaction
with cyclohexanol (TS2) with a barrier of 54.5 kJ mol-1 to form the with the additional pyridine ligand) but are lower in the case of the
alkyl peroxide chromium(VI) complex 5, which is exergonic by dialkyl peroxide (74.1 kJ mol-1, see the SI).
23.1 kJ mol-1. Addition via water (TS3) is also feasible with an We also looked at dichromic acid and the chromate and
activation energy of 75.1 kJ mol-1 and even direct addition without dichromate anions as potential catalysts, which could be
proton relay agents is possible (TS4, 86.2 kJ mol-1). Note, that the generated upon deprotonation through pyridine. The mechanistic
direct addition has a similar energetic barrier as the following picture is quite similar but the anions exhibit lower barriers for
deperoxidation step (TS6/TS7/TS8) so that even without proton dehydroperoxidation (79.0 kJ mol-1 for chromate, 71.5 kJ mol-1 for
relay agents, the overall barrier of the reaction would not change. dichromate, see SI), driven by the higher basicity of the oxo group
Under the catalytic conditions however, and regarding the in the anions. Overall, the discussed mechanism should be
generation of water from the reaction itself, we suggest that proton feasible for both monomeric chromic acid as well as its dimer and
relays are taking place. Since intermediates 1 and 5 are in the respective deprotonated anions. However, as shown in our
equilibrium, a large excess of water prevents formation of the experiments with dichromates we observed a drop in yield and
alkylperoxo intermediate 5 and is therefore disadvantageous. selectivity, which we attribute to side reactions only possible with
Similarly, an addition can occur to monosubstituted chromic the dinuclear catalyst while the regular chromic acid/chromate
acid derivatives or directly to chromium trioxide (see SI), albeit gives clean conversion to the ketone.
being higher in energy. Intermediate 5 can then go through the We also investigated a mechanism proposed by Boitsov et al.,
six-membered transition state TS6 with a barrier of 83.7 kJ mol-1, who suggested a mechanism on the basis of NMR experiments,
in which the hydrogen of the peroxide carbon is transferred to an the ketone is formed from the corresponding alcohol via an
accepting oxo group on the chromium(VI) center, while the O-O alkoxide/alkyl peroxide complex (see Figure 10).[51]
bond is cleaved to release cyclohexanone.

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
mechanistic studies including DFT calculations revealed that the
∆G chromium(VI)-catalyzed decomposition of cyclohexyl
O
kJ/mol
O
hydroperoxide happens via the formation of a
Cr O
O cyclohexylperoxychromium intermediate. These observations are
O O Cr O
O
O
H O in agreement with our previous work using different vanadium(V)
H O H complexes as catalysts.[21] The initial hypothesis of the
TS10
H
importance of a terminal M=O group[21] for a non-radical and
TS11
therefore selective dehydroperoxidation was confirmed using
(258.4 kJ/mol)
CrO3 as catalyst in this work. Previous studies using chromium(VI)
as catalyst presented different mechanisms regarding the
selective decomposition of cyclohexyl hydroperoxide.[41,51] A
comparison of the soluble and insoluble catalysts with respect to
their kinetic behavior showed that both systems follow the same

Accepted Manuscript
mechanism.
OH O
O The DFT calculations performed on the chromium-pyridine
+ +
O
Cr
O complex are supporting that the selective decomposition of
O
cyclohexyl hydroperoxide proceeds via the expected non-radical,
Cr O non-redox mechanism through formation of the
O O
O
cyclohexylperoxychromium complex as the reactive intermediate.
9 Furthermore, the experimentally obtained activation parameters
are in good agreement with the calculated values, indicating that
Figure 10. “Cooperative” alcohol/hydroperoxide mechanism suggested by the proposed mechanism is feasible.
Boitsov et al.[51] In this mechanism, the corresponding alcohol is merely an Due to the increasing demand for polyamides,[62] the development
intermediate resulting in the same net reaction. The process is however of efficient as well as cheap catalysts for the key step, the
energetically far less feasible than the mechanism shown in Figure 9
(168.9 kJ mol-1 vs. 83.7 kJ mol-1). formation of cyclohexanone, is of current interest in industry. As
presented, the use of insoluble chromium(VI) catalysts provides
Formation of chromium trioxide, cyclohexanol and cyclohexanone an alternative way to reduce the steps in the oxidation of
from the cyclohexanol cyclohexyl hydroperoxide diester of cyclohexane to cyclohexanone. Furthermore, environmental risks
chromic acid (9) is exergonic by 22.3 kJ mol-1, but the barrier for due to leaching of toxic Cr(VI) ions could be excluded. Finally, the
this reaction is 168.9 kJ mol-1 via TS10 (the proton relay through presented heterogeneous polymer-supported chromium(VI) is a
water, TS11, is even higher in energy with 258.4 kJ mol-1), making cheap as well as reusable catalyst and seems to be a promising
it energetically not feasible. Several effects likely play into this: the system for an industrial application.
lower basicity of the peroxide oxygen to accept the incoming
proton from the alkoxide, the weaker acidity of the proton on the
alkoxide relative to the hydrogen on the peroxide and the
relatively strong Cr-O bond, which needs to be broken, while the Experimental Section
previous mechanism only cleaves the weak O-O bond.
General Methods. Unless otherwise specified, reactions were carried out
under air in sealed NMR tubes or round bottom flasks. NMR spectra were
recorded using Bruker 200 at CaRLa or Bruker AVANCE III 300, Bruker
Conclusions
AVANCE III 400, Bruker AVANCE III 500 and Bruker AVANCE III 600
spectrometers at the Institute of Organic Chemistry / Heidelberg University.
Well-known homogeneous and heterogeneous chromium(VI) Chemical shifts are given in ppm referenced to the solvent (1H, 13C). IR
catalysts were investigated regarding their behavior in the vibrational spectra were recorded from 3500 to 400 cm-1 (Varian 2000,
selective decomposition of cyclohexyl hydroperoxide towards Scimitar Series, FTS2000) as KBr pellets at room temperature. In situ (low
cyclohexanone. This reaction plays an important role in the temperature and room temperature measurements) IR spectra were
industrial cyclohexane oxidation and an understanding of the recorded from 2000 to 650 cm-1 using a Mettler Toledo React IR 10. UV/vis
mechanism provides the opportunity to use very efficient measurements (low temperature and room temperature) were performed
using an Agilent 8453 spectrophotometer with a Unisoku USP-203-A
chromium(VI) catalysts for this reaction. With this study, the
Cryostat. The measurements were performed in solution in quartz glass
mechanism behind the chromium(VI) catalyzed
cuvettes, l = 1 cm. The chromium contents were determined in the central
dehydroperoxidation reaction was clarified. The chosen catalysts analytical laboratory of BASF SE in Ludwigshafen using ICP-MS.
CrO3*py and CrPVP were used in dehydroperoxidation reactions
under the same conditions and it was found that both systems Materials. Unless otherwise specified, reagents and solvents were
have a similar behavior regarding their catalytic activities and purchased from commercial suppliers (Aldrich, abcr, TCI) and used without
selectivities towards cyclohexanone. Especially at higher further purification. All chemicals were used as received.
temperatures, relevant for possible industrial applications, the
chromium(VI) catalysts show high catalytic activity with almost no Computational details. All structures were optimized at the BP86/def2-
decrease in the yield of the desired product. The comprehensive SV(P) level,[63-65] with final energies computed at the PBE0-D3(BJ)/def2-

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
QZVPP level of theory.[66] For all electronic structure calculations, the chloroform or cyclohexane, which are used as purchased and handled
TURBOMOLE program[67] was used. Zero-point vibrational energies and under air. The reactions were carried out in sealed NMR tubes at different
thermal corrections for the gas phase were obtained at the level of temperatures. The ratios of ketone to alcohol and the conversion of CHHP
geometry optimization and 298.15 K / 1 bar with TURBOMOLE’s freeh were measured by integration of the corresponding characteristic signals
program. Free enthalpies of solvation at infinite dilution in cyclohexane in 1H NMR spectra and naphthalene was used as internal standard. In a
were computed with COSMO-RS theory[68] using the COSMOtherm general procedure, a NMR tube was charged with the catalyst (1 eq.), the
program (Version C3.0, Release 1501, revision 1744).[69] substrate cyclohexyl hydroperoxide (150-1500 eq.), naphthalene and
deuterated solvent (0.5 mL) at room temperature. The samples were
Preparation of the catalysts. Preparation of CrO3*py. CrO3 (4.66 mg, stored at room temperature or heated up and the reactions were followed
46.6 µmol) was dissolved in pyridine-d5 (5.00 mL). The light-yellow by NMR measurements.
solution was stored at 5 °C.
In situ UV/vis measurements. Catalytic reaction. For following the
Preparation of CrPVP. The CrPVP catalysts were prepared according to decomposition reaction, a solution of CrO3*py in cyclohexane (0.93 mM)
the procedure reported by Lončarevic et al.[35] CrO3 (1.00 g, 10.0 mmol) was prepared. After measuring an UV/vis spectrum of the pure catalyst,
cyclohexyl hydroperoxide (500 eq.) was added and the rapid

Accepted Manuscript
was added in small portions to a suspension of PVP (10.0 g) in H2O
(100 mL). The light-yellow suspension was stirred overnight at room decomposition was followed for 60 min recording an absorption spectrum
temperature. After filtration, the yellow powder was washed several times every 10 seconds. Stoichiometric reaction. For following the formation of
with H2O until the filtrate became colorless. The catalyst was dried in the reactive intermediate, a solution of CrO3*py in α,α,α-trifluorotoluene
vacuum. The content of chromium on PVP was 2.6 wt. % determined by (0.93 mM) was prepared. After measuring an UV/vis spectrum of the pure
ICP-MS. The CrPVP catalyst with a higher or lower chromium loading were catalyst at 5 °C, cyclohexyl hydroperoxide (1 eq.) was added and the rapid
prepared in a similar manner. formation of the intermediate was followed for 2 min recording an
Preparation of MePVP. Poly(4-vinylpyridine) (10.0 g) was suspended in absorption spectrum every 10 seconds at 5 °C.
ethylene carbonate (35.0 mL). The mixture was stirred at 35 °C for 1 h.
Methyl methanesulfonate (12.0 mL, 0.142 mmol) was added. The mixture Low temperature NMR measurements. For following the formation of the
was heated to 90 °C for 24 h. Additional methyl methanesulfonate (4.0 mL, reactive intermediate, a solution of cyclohexyl hydroperoxide (0.8 mM) in
0.05 mmol) was added and the mixture was heated to 90 °C for further toluene-d8 was prepared. After measuring a 1H NMR spectrum of the pure
48 h. After filtration, the solid was washed with cold H2O and diethyl ether substrate at -20 °C, CrO3*py (1 eq.) was added and the rapid formation of
several times. The beige solid was dried in vacuum. the intermediate was followed for 3 h recording a NMR spectrum 10 min
Preparation of CrMePVP. The CrMePVP catalyst was prepared according at -20 °C.
to a procedure described by Tamami et al.[59] K2Cr2O7 (1.75 g, 5.95 mmol)
was added in small portions to a suspension of MePVP (2.00 g) in H2O In situ IR measurements. For following the formation of the reactive
(50 mL). The yellow suspension was stirred overnight at room temperature. intermediate, cyclohexyl hydroperoxide (250 eq.) was charged in a flask
After filtration, the dark yellow powder was washed several times with H2O and cooled to -74 °C. CrO3 (1 eq.) was added and the reaction was
until the filtrate became colorless. The catalyst was dried in vacuum. The followed recording an IR spectrum every 15 seconds. During the reaction,
content of chromium on MePVP was 3.2 wt. % determined by ICP-MS. the mixture was slowly warmed to room temperature and the formation of
cyclohexanone was followed for 3.5 h.
Preparation of cyclohexyl hydroperoxide. Pure cyclohexyl
hydroperoxide was isolated from the crude KA oil mixture (2 L, containing Kinetic measurements. For running temperature depending NMR
> 95 % cyclohexane, < 5 % cyclohexanol, cyclohexanone and cyclohexyl measurements, the catalytic decomposition of cyclohexyl hydroperoxide
hydroperoxide) obtained from the plant at BASF SE, Ludwigshafen. The was followed at 50 °C, 58 °C and 66 °C. A NMR tube was charged with
KA oil mixture was cooled to 5 °C and sodium hydroxide (40 % solution, cyclohexyl hydroperoxide (400 eq.) and naphthalene followed by the
500 mL) was slowly added. Caution! The product of this step can be addition of 1 eq. catalyst. The catalytic reaction was monitored for 80 min.
explosive! The sodium salt was filtered and washed with acetone (3 x 400 The ratios of ketone to alcohol and the conversion of CHHP were
mL) and cyclohexane (400 mL). The slightly yellow solid was dried in measured by integration of the corresponding characteristic signals in the
vacuum. The solid was then dissolved in water (800 mL) and cyclohexane 1
H NMR spectra using naphthalene as internal standard.
(400 mL). During cooling with ice, phosphoric acid (20 % solution) was
added until pH 7 was reached. The formed solid was washed with
cyclohexane (2 x 100 mL). The organic phase was dried over sodium
sulfate and evaporated under reduced pressure. The obtained yellowish Acknowledgements
oil was purified by column chromatography (hexane: ethyl acetate, 10:1).
Cyclohexyl hydroperoxide could be obtained as a clear, colorless oil. Yield: CaRLa (Catalysis Research Laboratory) is co-financed by the
(8.42 g, 72.5 mmol). Rf = 0.23. 1H NMR (200 MHz, CDCl3): δ = 7.67 (s, 1H, Heidelberg University and BASF SE.
OH), 3.96 (m, 1H, CH), 1.96 – 1.28 (m, 10H, CH2) ppm. 13C NMR (50 MHz,
CDCl3): δ = 83.4 (1C, CH), 30.2 (2C, CH2), 25.8 (1C, CH2), 23.8 (2C, CH2)
ppm.
Conflict of interest
The NMR data of the isolated cyclohexyl hydroperoxide are in perfect
agreement with the synthesized cyclohexyl hydroperoxide using a The authors declare no conflict of interest.
procedure according to Kundu et al.[70] In the catalytic decompositions,
both cyclohexyl hydroperoxides show an identical behavior.

Dehydroperoxidation of cyclohexyl hydroperoxide. Each catalytic


Keywords: Cyclohexane oxidation • cyclohexyl
reaction was carried out several times, and herein, the results are hydroperoxide • chromium • homogeneous catalysis •
presented as average values. The reactions were performed in deuterated kinetic

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
[1] I. C. Twilley, D. W. H. Roth, Jr., R. A. Lofquist, US3558567 (A), 1971. [44] D. Lončarevic, J. Krstic, P. Bankovic, S. Anic, Z. Čupić, Russian Journal
[2] L. L. Estes, M. Schweizer, Fibers, 4. Polyamide Fibers. In Ullmann's of Physical Chemistry, A 2007, 81, 1398-1401.
Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2000. [45] D. Lončarevic, Z. Čupić, Materials Science Forum 2005, 494, 363-368.
[3] H.-J. Arpe, Industrial Organic Chemistry, 5th ed., Wiley-VCH, Weinheim, [46] K. W. Jun, E. K. Shim, S. E. Park, K. W. Lee, Bull. Korean Chem. Soc.
2010, p. 243. 1995, 16, 398-400.
[4] J. Ritz, H. Fuchs, H. Kieczka, W. C. Moran, Caprolactam. In Ullmann's [47] D. Lončarevic, Z. Čupić, M. Odović, J. Serb. Chem. Soc. 2005, 70, 209-
Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2011. 221.
[5] A. Kuessner, G. Herrmann, US 3917708, 1975. [48] T. Brunelet, G. Gelbard, A.Guyot, Polymer Bulletin 1981, 5, 145-149.
[6] K. Pugi, US 3530185, 1970. [49] L. Li, Z. Wu, L. Guo, X. Yuan, Can. J. Chem. Eng. 2016, 94, 1987-1994.
[7] J. Hermolin, US 4465861, 1984. [50] B. M. Weckhuysen, I. E. Wachs, R. A. Schoonheydt, Chem. Rev. 1996,
[8] J. H. Teles, I. Hermans, G. Franz, R. A. Sheldon, Oxidation. In Ullmann's 96, 3327-3350.
Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2000. [51] S. Boitsov, J. Songstad J. Muzart, J. Chem. Soc., Perkin Trans. 2 2001,
[9] J. D. Druliner, S. D. Ittel, P. J. Krusic, C. A. Tolman, US 4326084, 1982. 2318-2323.
[10] M. T. Musser, Cyclohexanol and Cyclohexanone. In Ullmann's [52] a) A. D. Becke, Phys. Rev. A 1988, 38, 3098-3100; b) C. Lee, W. Yang,
Encyclopedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2000. R. Parr, Phys. Rev. B 1988, 37, 785-789; c) A. D. Becke, J. Chem. Phys.

Accepted Manuscript
[11] W. J. Barnette, D. L. Schmitt, J. O. White, US 3987100, 1976. 1993, 98, 5648-5652.
[12] M. Costantini, E. Fache, DE 69815580, 2004. [53] J. M. J. Frechet, M. Vivas de Meftahi, Brit. Poly. J. 1984, 16, 193-198.
[13] D. L. Valdez, WO/2014/014467, 2014. [54] H. Firouzabadi, B. Tamami, N. Goudarzian, M. M. Lakouraj, M. Hatam,
[14] M. Rapoport, J. O. White, US 3957876, 1976. Synth. Commun. 1991, 21, 2077.
[15] P. V. Saji, C. Ratnasamy, S. Gopinathan, US 6392093, 2002. [55] J. M. J. Frechet, J. Warnock, M. J. Farrall, J. Org. Chem. 1978, 43, 2618-
[16] R. Raja, P. Ratnasamy, Catal. Lett. 1997, 48, 1-10. 2621.
[17] M. Dugal, G. Sankar, R. Raja, J.M. Thomas, Angew. Chem., Int. Ed. [56] J. M. J. Frechet, P. Darling, M. J. Farrall, J. Org. Chem. 1981, 46, 1728-
2000, 39, 2310-2313. 1730.
[18] R. Raja, G. Sankar, J. M. Thomas, J. Am. Chem. Soc. 1999, 121, 11926- [57] E. A. Boucher, C. C. Mollett, J. Chem. Soc., Faraday Trans. 1 1982, 78,
11927. 75-88.
[19] I. W. C. E. Arends, R. A. Sheldon, Appl. Catal. A 2001, 212, 175-187. [58] L. Gargallo, B. Miranda, H. Ríos, F. González-Nilo, D. Radić, Polym. Int.
[20] C. A. Tolman, J. D. Druliner, P. J. Krusic, M. J. Nappa, W. C. Seidel, I. D. 2001, 50, 858-862.
Williams, S. D. Ittel, J. Mol. Catal. 1988, 48, 129-148. [59] B. Tamami, H. Firouzabadi, M. Mansour Lakouraj, A.-R. Mahdavian, Iran.
[21] A.-C. Schmidt, M. Hermsen, F. Rominger, R. Dehn, J. H. Teles, A. J. Polymer Sci. Technol. 1994, 3, 82-87.
Schäfer, O. Trapp, T. Schaub, Inorg. Chem. 2017, 56, 1319-1332. [60] T. Schaub, P. Fischer, A. Steffen, T. Braun, U. Radius, A. Mix, J. Am.
[22] R. R. Hiatt, K. C. Irwin, C. W. Gould, J. Org. Chem. 1968, 33, 1430-1435. Chem. Soc. 2008, 130, 9304-9317.
[23] J. E. Lyons, P. E. Ellis, Jr., US5120886, 1992. [61] J. C. Collins, W. W. Hess, F. J. Frank, Tetrahedron Lett. 1968, 9, 3363-
[24] R. A. Sheldon, J. K. Kochi, Academic Press 1981, 33-70. 3366.
[25] W. Pritzkow, K. A. Müller, Chem. Ber. 1956, 89, 2321-2328. [62] H.-G. Elias, R. Mülhaupt, Plastics, General Survey, 1. Definition,
[26] R. R. Hiatt, T. Mill, K. C. Irwin, J. K. Castleman, J. Org. Chem. 1968, 33, Molecular Structure and Properties. In Ullmann's Encyclopedia of
1428-1430. Industrial Chemistry, Wiley-VCH, Weinheim, 2000.
[27] C. B. Hansen, G. J. Hoogers, W. Drenth, J. Mol. Catal. 1993, 79, 153- [63] J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33,
163. 8822-8824.
[28] G. F. Pustarnakova, V. M. Solyanikov, E. T. Denisov, Izv. Akad. Nauk [64] A. D. Becke, Phys Rev A Gen Phys 1988, 38, 3098-3100.
SSSR, Ser. Khim. 1975, 547-552. [65] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297-3305.
[29] E. T. Denisov, A. F. Shestakov, Russ. Chem. Bull. 2012, 61, 17-27. [66] M. Ernzerhof, G. E. Scuseria, J. Chem. Phys. 1999, 110, 5029-5036.
[30] J. D. Chen, J. Dakka, R. A. Sheldon, Appl. Catal., A 1994, 108, L1-L6. [67] R. Ahlrichs, M. K. Armbruster, R. A. Bachorz, M. Bär, H.-P. Baron, R.
[31] L. V. Petrov, V. M. Solyanikov, B. Acad. Sci. USSR CH+ 1991, 40, 1958- Bauernschmitt, F. A. Bischoff, S. Böcker, A. M. Burow, N. Crawford, P.
1960. Deglmann, F. Della Sala, M. Diedenhofen, M. Ehrig, K. Eichkorn, S.
[32] L. Saussine, E. Brazi, A. Robine, H. Mimoun, J. Fischer, R. Weiss, J. Am. Elliott, D. Friese, F. Furche, A. Glöß, N. Graf, L. Grajciar, F. Haase, M.
Chem. Soc. 1985, 107, 3534-3540. Häser, C. Hättig, A. Hellweg, B. Helmich, S. Höfener, H. Horn, C. Huber,
[33] F. A.Chavez, J. M. Rowland, M. M. Olmstead, P. K. Mascharak, J. Am. U. Huniar, M. Kattannek, W. Klopper, A. Köhn, C. Kölmel, M. Kollwitz, M.
Chem. Soc. 1998, 120, 9015-9027. Kühn, K. May, P. Nava, C. Ochsenfeld, H. Öhm, M. Pabst, H. Patzelt, D.
[34] R. Hofmann, H. Hübner, G. Just, L. Krätzsch, A. K. Litkowez, W. Pritzkow, Rappoport, O. Rubner, A. Schäfer, G. Schmitz, U. Schneider, M. Sierka,
W. Rolle, M. Wahren, J. Prakt. Chem. 1968, 37, 102-112. D. P. Tew, O. Treutler, B. Unterreiner, M. v. Arnim, F. Weigend, P. Weis,
[35] D. Lončarevic, J. Krstic, J. Dostanić, D. Manojlovic, Z. Čupić, D. M. H. Weiss, Nina Winter, TURBOMOLE V7.0; TURBOMOLE GmbH:
Jovanović, Chem. Eng. J. 2010, 157, 181-188. Karlsruhe, Germany, 2015. http://www.turbomole.com .
[36] M. Wang, J. Ma, C. Chen, X. Zheng, Z. Du, J. Xu, J. Mater. Chem. 2011, [68] F. Eckert, A. Klamt, AlChE J. 2002, 48, 369-385.
21, 12609-12612. [69] COSMOtherm, Version C3.0, Release 1501; COSMOlogic GmbH & Co
[37] A. Bréhéret, C. Lambeaux, S. Ménage, M. Fontecave, F. Dallemer, E. KG: Leverkusen, Germany, 2014. http://www.cosmologic.de.
Fache, J.-P. Pierre, P. Chautemps, M.-T. Averbusch-Pouchot, C. R. [70] R. Kundu, Z. T. Ball, Org. Lett. 2010, 12, 2460-2463.
Acad. Sci. Paris, Série IIc, Chimie 2001, 4, 27-34.
[38] J. Muzart, Tetrahedron Lett. 1987, 28, 2133-2134.
[39] J. J. M. Va Donck, J. L. J. P. Hennekens, W. Voskuil, J. Wolters, DE
2352378, 1973.
[40] S. Hamamoto, K. Harada, M. Hayashi, JP 58072532, 1983.
[41] W. Buijs, R. Raja, J. M. Thomas, H. Wolters, Catal. Lett. 2003, 91, 253-
259.
[42] S. Chouzier, S. Veracini, F. Igersheim, WO 2011064135 (A1), 2010.
[43] M. Mansour Lakouraj, A. Keyvan, J. Chem. Research 1999, 206-207.

This article is protected by copyright. All rights reserved.


ChemCatChem 10.1002/cctc.201701909

FULL PAPER
Entry for the Table of Contents (Please choose one layout)
Layout 1:

FULL PAPER
A look behind the scenes: Known Jessica Nadine Hamann, Marko
homogeneous and heterogeneous Hermsen, Anna-Corina Schmidt, Saskia
Cr(VI) systems were used to Krieg, Jasmin Schießl, Dominic Riedel,
investigate the mechanism behind the Joaquim Henrique Teles, Ansgar

Accepted Manuscript
transition metal-catalyzed selective Schäfer, Peter Comba, A. Stephen K.
decomposition of cyclohexyl Hashmi, Thomas Schaub*
hydroperoxide. Based on
comprehensive experimental as well Page No. – Page No.
as DFT studies, a non-radical, non-
Selective decomposition of
redox mechanism via a
cyclohexyl hydroperoxide using
cyclohexylperoxychromium(VI)
homogeneous and heterogeneous
complex is proposed.
Cr(VI) catalysts: Optimizing the
reaction by evaluating the reaction
mechanism

This article is protected by copyright. All rights reserved.

You might also like