Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Catalysis

Science & Technology


View Article Online
PAPER View Journal | View Issue

Hydrogenation of acrylonitrile–butadiene copolymer


Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

Cite this: Catal. Sci. Technol., 2013,


latex using water-soluble rhodium catalysts
3, 2689
Yin Liu, Hanmiroo Kim, Qinmin Pan* and Garry L. Rempel*

Hydrogenated acrylonitrile butadiene rubber (HNBR) is a high-performance elastomer which has found
important applications in the automobile and petroleum industries. Compared to current NBR solution
hydrogenation technology, direct hydrogenation of NBR in its aqueous form (i.e. NBR latex) is a
‘‘green’’ strategy as it is energetically favorable and environmentally friendly. Two water-soluble
rhodium based catalysts, RhCl(TPPMS)3 (TPPMS = monosulfonated triphenylphosphane) and
RhCl(TPPTS)3 (TPPTS = trisulfonated triphenylphosphane), were investigated for NBR latex
hydrogenation. The NBR in aqueous latex could be hydrogenated using the RhCl(TPPMS)3 catalyst,
attaining HNBR with as much as 95 mol% CQC bond removal. A limited fraction of CQC bond
reduction was observed when RhCl(TPPTS)3 was used as the catalyst. Using the RhCl(TPPMS)3 catalyst,
the hydrogenation reaction could be conducted at mild temperature with no addition of organic
co-solvent. The presence of Rh metal in the resultant HNBR confirmed the phase transfer of the
RhCl(TPPMS)3 catalyst and was considered as a crucial factor for achieving a successful reaction. The
observed influence of process conditions on the activity of hydrogenation has led to an extrapolation
Received 16th April 2013, of a mild reaction condition mechanism for the RhCl(TPPMS)3 system. Additionally, in-house NBR latices
Accepted 2nd July 2013 with different gel content were synthesized and investigated in the hydrogenation experiment for
DOI: 10.1039/c3cy00257h comparison. One of the major observations is that the high gel fraction in NBR could greatly limit the
hydrogenation reaction which could be a reasonable explanation for the low conversion of CQC bonds
www.rsc.org/catalysis reported previously.

1. Introduction carbon–carbon double bonds (CQC bonds) in NBR were


reduced while the unsaturated carbon–nitrogen triple bonds
Hydrogenation of unsaturated diene-based polymers constitutes (CRN bonds) remained unchanged (Fig. 1). The high manu-
an important process in polymer chemistry because of the desir- facturing cost of HNBR caused by the cumbersome hydrogena-
able properties of the materials obtained.1,2 The hydrogenation of tion procedures such as obtaining solid NBR from NBR latex,
acrylonitrile butadiene rubber (NBR) to produce hydrogenated-
NBR (HNBR) represents an excellent example since (1) HNBR has
improved chemical, physical and mechanical properties com-
pared to NBR. It imparts excellent resistance to oxidative and
ozonolytic ageing, improved resistance to oils and fluids at high
temperatures and reduced gas permeability; (2) the ethylene–
acrylonitrile co-polymer, which has a similar chemical structure
to HNBR, is difficult to synthesize by conventional polymerization
(i.e. low degree of polymerization).
Currently, most HNBR is manufactured via a solution hydro-
genation process where NBR was dissolved in a certain organic
solvent and hydrogenation was achieved by using H2 gas in the
presence of a precious metal based catalyst (e.g. Pd based supported
catalysts and Rh based organometallics).3 During the reaction, the

University of Waterloo, Waterloo, Ontario, Canada N2L 3G1.


E-mail: qpan.uw@gmail.com, grempel@uwaterloo.ca Fig. 1 Schematic of NBR hydrogenation.

This journal is c The Royal Society of Chemistry 2013 Catal. Sci. Technol., 2013, 3, 2689--2698 2689
View Article Online

Paper Catalysis Science & Technology

dissolving solid NBR in an organic solvent and separating water soluble rhodium based complexes modified with the
solvent/catalyst from the HNBR product greatly limits its ligands PPh2(CH2)n–COONa (n = 5, 7).9 Similar low conversions
applications. (50–60%) were observed for all three polymers and massive
NBR is synthesized via emulsion polymerization and can be leaching of the rhodium from the aqueous phase to the organic
kept as NBR latex. Direct hydrogenation of NBR in latex form phase was observed.
without using an organic solvent would be more economical Recently, Papadogianakis and co-workers demonstrated a
and environmentally friendly. It has been recommended as a successful catalytic hydrogenation of an unsaturated polymer
potential ‘‘green’’ pathway to produce HNBR. Hence, direct polybutadiene-1,4-block-poly(ethylene oxide) (PB-b-PEO) in
hydrogenation of NBR latex has attracted attention from both mixed micellar nano-reactors formed by dodecyltrimethyl-
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

industry and academia where considerable efforts have been ammonium chloride (DTAC) using Rh–TPPTS complexes in
made to realize this process. Research on NBR latex hydrogena- aqueous media.24 An exceptionally high degree of hydrogena-
tion has focused on either a hydrazine-diimide route (i.e. tion (100%) and activity (TON > 840 h1) was observed. In
utilization of hydrazine in the presence of hydrogen peroxide addition, its catalytic activity was observed even at a rhodium
to produce diimide in situ for CQC bonds reduction)4–7 or a concentration of only 1 ppm in water.
gaseous hydrogen route (i.e. use of hydrogen gas for reduction In our lab, we observed that RhCl(PPh3)3 catalyst is active for
with a hydrogenation catalyst).8–12 The first HNBR latex has hydrogenation of unsaturated polymers in aqueous systems
been commercialized by Zeon Chemical under the trademark (e.g. NBR, SBR latex) with the addition of extra triphenylphos-
Chemisat in 2008 via the hydrazine-diimide route.13 Currently, phine (PPh3).11,25 This important finding evokes our interest to
the core challenges in catalytic hydrogenation of NBR in latex re-investigate the polymer latex hydrogenation using its water
form lie in areas including (1) exploration of active catalysts to soluble analog catalysts. In this research work, we report the
achieve desired high conversion, (2) avoidance of the use of an hydrogenation of a typical unsaturated polymer, NBR, in its
organic co-solvent, (3) control of cross-linking side reaction, latex form using two common water-soluble rhodium catalysts,
and (4) recovery of the catalyst. RhCl(TPPMS)3 and RhCl(TPPTS)3, in the absence of any added
Catalytic hydrogenations in aqueous media have been organic solvent. Very different hydrogenation behaviors were
reported since the 1960s.14,15 In the last two decades, increasing observed when using these two water soluble catalysts.
interest in catalysis in aqueous media has been pursued
directly by employing water-soluble transition metal com- 2. Experimental
plexes.16 Water-soluble catalysts could intensify the catalytic
reaction by averting the mass transfer limitation in water which 2.1 Materials
is commonly encountered when using conventional organo- The catalysts, RhCl(TPPMS)3 and RhCl(TPPTS)3, were prepared
metallics. The use of water soluble rhodium catalysts such as by reacting RhCl3 with the corresponding ligand (TPPMS, TPPTS)
RhCl(TPPMS)3 [TPPMS: monosulfonated triphenylphosphine, in ethanol.26,27 Reagent grade mono-chlorobenzene (MCB) was
PPh2(C6H4-m-SO3Na)], Rh(TPPTS)3 [TPPTS: trisulfonated triphenyl- obtained from Fisher Scientific (Canada). Reagent grade toluene,
phosphine, P(C6H4-m-SO3Na)3] and RhCl(PTA)3 [PTA: 1,3,5- ketones (e.g. acetone, 2-butanone) and alcohols (e.g. methanol
triaza-7-phosphaadamantane, C6H12N3P] has been the subject and ethanol) were obtained from Sigma-Aldrich (Canada). All of
of research for many years.17,18 Some of these catalysts have the solvents were used as received.
provided significant achievements in the hydroformylation of Two types of NBR latices were used in this investigation. The
small olefins.19,20 commercial NBR latex was provided by LANXESS Deutschland
With the success of the utilization of water soluble catalysts GmbH. The in-house NBR latex was synthesized in the lab
for small olefin hydrogenation, these catalysts have also been following a standard emulsion polymerization procedure.28
studied when large olefins (e.g. C6–C10) are used as reac-
tants.21,22 However, the catalyst showed less activity towards 2.2 Synthesis of NBR via emulsion polymerization
large olefins in the biphasic system. The activity of the catalysts Reagent grade acrylonitrile (99%), potassium persulfate, tert-
becomes even worse when they are applied for poly-olefin dodecyl mercaptan (mixture of isomers, 98.5%), hydroxylamine
substrates. Specifically, the conversion (i.e. the percentage of solution (50 wt% in H2O) and technical grade sodium oleate
CQC bonds hydrogenated to C–C bonds) is insufficient to meet (83 wt%) were used as received from Sigma-Aldrich (Canada).
industrial requirements. Singha et al. first studied the hydro- Technical grade 1,3-butadiene (99%) was obtained from Air
genation of a commercial NBR latex using water soluble Liquide (Canada).
RhCl(TPPMS)3 catalysts with the addition of the non-ionic NBR latex was synthesized via emulsion copolymerization.
surfactant Triton-X305.23 They found that only 60% CQC The polymerization reaction was performed in a 300 mL stain-
bonds in NBR could be hydrogenated with a TON (turnover less steel autoclave (Parr Instrument, USA). The reactor, chilled
number, calculated based on the final conversion) between in an ice–water bath, was charged with distilled water, acrylo-
4.0 and 9.3 h1. After that, Mudalige and Rempel investigated nitrile (monomer), sodium oleate (surfactant), tert-dodecyl mer-
the hydrogenation of polybutadiene (Mn = 900 g mol1), NBR captan (chain transfer agent) and potassium persulfate (initiator).
(Mn = 6400 g mol1) and styrene–butadiene rubber (Mn = After assembling the reactor, the mixture was degassed and purged
95 000 g mol1) in an aqueous–organic biphasic system using with nitrogen gas for 30 minutes. Then liquefied 1,3-budadiene

2690 Catal. Sci. Technol., 2013, 3, 2689--2698 This journal is c The Royal Society of Chemistry 2013
View Article Online

Catalysis Science & Technology Paper

(monomer) was quickly added and copolymerized with acrylo- sampled at intervals to measure its hydrogenation degree (hydro-
nitrile at 30 1C for several hours. After reaching a desired genation degree is defined in Section 3.1).
conversion, hydroxylamine was added to terminate the reac-
tion. The residual portions of acrylonitrile and 1,3-butadiene
were eliminated by steam distillation at 70 1C under reduced 2.4 Characterization
pressure. The hydrogenation degree in the resultant product was
obtained via 1H NMR. Room temperature 1H NMR spectra of
NBR/HNBR in CDCl3 were taken on a Bruker 300 MHz Spectro-
2.3 NBR latex hydrogenation meter (Bruker BioSpin Corp., Massachusetts, USA) with TMS as
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

The hydrogenation reaction was carried out in a 300 mL high internal standard.
pressure reactor (Parr Instrument, USA). 100 mL NBR latex Fig. 2 shows the chemical shifts of the protons in (a) NBR
(or with a certain amount of additives such as organic co-solvents, and (b) HNBR by 1H NMR characterization. For NBR, the peaks
co-catalyst ligands or surfactants) was added to the reactor. The from 4.9 to 5.6 ppm represent the olefinic protons. Among
catalyst RhCl(TPPMS)3 or RhCl(TPPTS)3 was weighed in a glass them, the large peaks between 5.3 and 5.6 ppm are related to
bucket and placed in a catalyst addition device which was trans hydrogen atoms of butadiene resulting from 1,4-addition
installed in the head of the reactor. After completely degassing and the small peaks between 4.9 and 5.1 ppm belong to the
the reactor with N2 gas, the catalyst was charged to the mixture hydrogen atoms of pendant vinyl groups which come from
and the reactor was heated to the desired temperature. The 1,2-addition. Meanwhile, all of the aliphatic protons in the CH3,
hydrogen pressure and reaction temperature were kept constant CH2 and CH microstructures lie between 0.8 and 2.8 ppm. In
throughout the reaction period. The hydrogenated NBR latex was HNBR (ca. 95 mol%), it could be clearly seen that most of the

Fig. 2 Typical 1H NMR spectra of (a) NBR and (b) hydrogenated NBR.

This journal is c The Royal Society of Chemistry 2013 Catal. Sci. Technol., 2013, 3, 2689--2698 2691
View Article Online

Paper Catalysis Science & Technology

peaks in the olefinic proton region had disappeared due to the same as the conversion of CQC bonds. Based on the molecular
reduction of unsaturated CQC bonds by H2. weight of the polymer we used, the average number of CQC
The concentration of rhodium present in the HNBR was bonds in a single polymer chain should be more than a
measured using ICP-AES, assuming that it comes from the thousand.
catalyst. Before being analyzed using ICP-ACE, the HNBR needs Second, strictly speaking, HNBR should be referred to
to be digested:29 400–500 mg of HNBR polymer solid was completely hydrogenated NBR (no CQC bonds left). In this
weighed out and mixed with hydrochloric acid (1 mL, 37.5 vol%), research work, the term ‘‘HNBR’’ is used for hydrogenated NBR
nitric acid (5 mL, 62 vol%) and aqueous hydrogen peroxide with at least 95 mol% hydrogenation degree. Usually, a small
(1 mL, 30 wt%). The mixture was digested in a High Pressure portion of residual CQC bonds should be left for sulfur curing
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

Asher (Anton Parr, Austria), under a high ‘‘organic program’’ systems to develop optimal properties that render the HNBR
(program condition, 300 1C and 130 bar for 3 hours). The resistant to hydrocarbons, oils and plasticizers. Still, achieving
rhodium metal concentration in the resultant solution was a high hydrogenation degree is crucial in the hydrogenation of
measured using inductively coupled plasma atomic emission polymeric materials.
spectroscopy (ICP-AES, TELEDYNE, LEEMAN Labs, Prodigy, Third, the hydrogenation degree data obtained from NBR
high dispersion ICP, USA). Commercial rhodium standard hydrogenation were in the absence of any mass transfer effect.
solution (1000 mg L1 Rh in HCl, Sigma-Aldrich) was used as The gas–liquid and liquid–solid (polymer particle) mass trans-
standard reference for ICP calibration.30 fer effects were eliminated by carrying out the reaction by
The micro-structure of NBR or HNBR was measured by varying the magnetic stirrer speed from 100 rpm to 1000 rpm
Fourier-transform infrared (FT-IR) analysis using a Bio-Rad and observing no change in the NBR conversion in the reaction
FTS 3000MX spectrometer. 16 times of uniform scanning with (temperature > 80 1C) above an agitation speed of 600 rpm. To
a resolution of 4 cm1 was conducted in all cases. To prepare be on the safe side, all of the runs were carried out at 1000 rpm.
samples for FT-IR, the NBR or HNBR has to be isolated from The diffusion of the catalyst inside the NBR particles could be
the latex and dissolved in 5% (wt/vol.) methyl ethyl ketone. The negligible since the size of NBR particles is very small (B70 nm,
polymer solution was cast onto a single NaCl crystal disc and measured by dynamic light scattering31). In our experiment, the
dried to form a thin polymer film (o1 mm). catalyst was still pre-mixed with the NBR latex at the reaction
The weight average molecular weight (Mw), the number temperature before adding H2 gas to the system. The ICP-AES
average molecular weight (Mn) and the polydispersity index results showed that the rhodium concentration in NBR particles
(PDI), which is equal to the ratio of the weight average to the did not undergo any obvious change during the hydrogenation.
number average molecular weight, were measured using a gel Initial experiments using commercial NBR latex to afford
permeation chromatograph (GPC, Wyatt Technology Corpora- HNBR latex in the presence of RhCl(TPPMS)3 or RhCl(TPPTS)3
tion, California, USA) equipped with an RI detector (Waters catalysts are presented in Tables 1 and 2 respectively. The
150-CV refractive index detector) and a multi-angle laser light catalysts behaved differently with respect to the hydrogenation
scattering instrument (DAWNs DSP-F Laser Photometer). The
dried NBR solid from the NBR latex was dissolved in THF and
Table 1 NBR latex hydrogenation using the RhCl(TPPMS)3 catalyst
filtered through a 45 mm syringe filter with a 250 nm GHP
(GH Polypro) membrane (Pall Corporation, New York, USA) and RhCl(TPPMS)3 NBR solid [CQCNBR]/[Cat.] Time HD
then 100 mL of the solution was injected into the GPC system Entry (mmol L1) content (g L1) (mol/mol) (hours) (mol%)
using THF as an eluent at a flow rate of 1.0 mL min1 at 30 1C. A 1 0.26 25 1100 12 95
polystyrene standard (Mn = 96 722 g mol1 and Mw = 98 251 g mol1) 2 0.52 50 1100 5 80
3 0.52 50 1100 8.5 95
was used for calibration. 4 1.04 50 550 6 95
5 1.04 100 1100 20 95
3. Results and discussion Commercial NBR latex, total volume = 100 mL, T = 100 1C, PH2 = 1000 psi
(69 bar).
3.1 NBR latex hydrogenation using different water-soluble
rhodium catalysts
First, as we mentioned in the Introduction section, NBR hydro- Table 2 NBR latex hydrogenation using the RhCl(TPPTS)3 catalyst
genation refers to the hydrogenation of CQC bonds in NBR. To
RhCl(TPPTS)3 NBR solid [CQCNBR]/[Cat.] Time HD
accurately describe this reaction process, the hydrogenation Entry (mmol L1) content (g L1) (mol/mol) (hours) (mol%)
degree (HD) is used in this research work and defined as
1 0.26 25 1100 20 20
follows: 2 0.52 50 1100 5 15
  3 0.52 50 1100 10 19
½C ¼ Ct 4 0.52 50 1100 13 21
HD ¼ 1   100% (1)
½C ¼ C0 5 0.52 50 1100 20 22
6 1.04 100 1100 10 15
where [CQC]t is the concentration of CQC bonds at reaction 7 1.04 50 550 14 19
time t and [CQC]0 is the initial double bond concentration. We Commercial NBR latex, total volume = 100 mL, T = 100 1C, PH2 =
can see that the hydrogenation degree is defined as being the 1000 psig (69 bar).

2692 Catal. Sci. Technol., 2013, 3, 2689--2698 This journal is c The Royal Society of Chemistry 2013
View Article Online

Catalysis Science & Technology Paper

degree. When using the RhCl(TPPMS)3 catalyst, the hydrogena-


tion degree continuously went up and a HNBR with a 95 mol%
HD could be achieved at the end of the reaction. The 1H NMR
spectrum (Section 2.4) showed that most of the CQC bonds
were saturated during the reaction.
In addition, the infrared analysis of the hydrogenated pro-
duct revealed that no detectable reduction of nitrile unsatura-
tion (no peaks appearance between 3000 cm1 and 3200 cm1)
when the RhCl(TPPMS)3 catalyst is used, suggesting that the oil
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

resistant property of the material is not compromised by the


hydrogenation process. The macro gel in the resultant HNBR
could be simply measured by re-dissolving the dried HNBR
into monochlorobenzene. The HNBR product was completely
soluble and no visible gel was observed. All the above results
showed that the RhCl(TPPMS)3 catalyst is a very active catalyst
Fig. 3 The change of rhodium concentration in NBR polymer particles through
for NBR latex hydrogenation.
hydrogenation. Commercial NBR latex = 50 g L1, Rh catalysts = 0.52 mmol L1,
Next, we evaluated the NBR latex hydrogenation using the total volume = 100 mL, T = 100 1C, PH2 = 1000 psig, square: conversion, circle:
RhCl(TPPTS)3 catalyst. It was observed that the hydrogenation rhodium concentration.
conversion increased slowly during the first several hours and
the extent of hydrogenation leveled off, showing only a marginal
increase thereafter. No desired high hydrogenation degree could be of poly-butadiene has confirmed that the TPPTS ligand based
obtained in the presence of the RhCl(TPPTS)3 catalyst. In addition, rhodium catalyst is prone to stay in the aqueous phase.34
the RhCl(TPPTS)3 catalyst remained in the aqueous phase through- However, the distribution of the RhCl(TPPMS)3 in the hydro-
out the reaction as observed by a reddish aqueous phase. genation system of the high molecular weight polymer in the
Although both catalysts are water soluble, the catalytic aqueous latex system is quite different from that of RhCl-
behaviors with respect to NBR latex hydrogenation are quite (TPPTS)3. More than half of the catalyst diffused into the
different. As we understand, the cause of this discrepancy could polymer particles under our reaction conditions. The catalyst
be attributed to the intrinsic properties of the catalyst ligand. in the polymer phase was extracted from the aqueous phase
Due to the inherent surfactant property of the TPPMS ligand, during the reaction. The result here was remarkably different
the RhCl(TPPMS)3 complex could be accumulated in a high from the hydrogenation of small organic molecules or the
concentration on the interface.32 Compared to TPPMS, TPPTS biphasic hydrogenation of polymers with RhCl(TPPMS)3 where
has a larger number of sulfonate groups which makes it much the catalyst always remained in the aqueous phase after the
more hydrophilic. The solubility of TPPTS in water is 180 g L1 reaction.9,35 Although both RhCl(TPPMS)3 and RhCl(TPPTS)3
which is more than ten times that of TPPMS (12 g L1 in water).33 water soluble metal complexes have been numerously reported
As a result, the RhCl(TPPTS)3 complex prefers to remain in the as effective catalysts in aqueous biphasic catalysis for small
aqueous phase so that only the CQC bonds on the NBR surface olefins, it is evident here that RhCl(TPPMS)3 is more active for
could be contacted and hydrogenated. Since HNBR could not hydrogenation of heavy molecular weight polymer latex such as
be achieved when using RhCl(TPPTS)3, the following discus- NBR latex. To our knowledge, this interesting phenomenon has
sion will mainly focus on hydrogenation results obtained using not been reported or explained previously. We believe that the
the RhCl(TPPMS)3 catalyst. The results obtained from RhCl- proper relative partition of the RhCl(TPPMS)3 catalyst between
(TPPTS)3 are merely used for comparison. water and the polymer substrate caused successful phase
transfer of the catalyst during the hydrogenation reaction. How-
3.2 The ‘‘location’’ of the catalyst in NBR latex hydrogenation ever, the attribution of the surfactant to the RhCl(TPPMS)3
The distribution of the catalyst is quantitatively measured in catalyst distribution should be minimal based on the hydro-
order to explain the different behaviors of these two catalysts. genation results using the RhCl(TPPTS)3 catalyst.
Compared to analyzing the catalyst concentration in the aqueous In addition, we believe that the fraction of catalyst in the
phase, the measurement of the catalyst in the polymer phase polymer phase should be responsible for the hydrogenation
gives more accurate information for the location of the catalyst. reaction, particularly for achieving a high hydrogenation
From Fig. 3, it can be seen that a small portion of the degree. Therefore, in this water–polymer two phase reaction,
rhodium was detected in the polymer phase after the reaction using the feed catalyst concentration to calculate TON without
when using the RhCl(TPPTS)3 catalyst. The tri-sulfonated PPh3 measuring its location would be less meaningful. It is also
ligand (TPPTS) based rhodium catalyst is completely soluble in worth noting that catalyst transfer is not encouraged in the
the aqueous phase. Since the catalyst mainly remains in the biphasic reaction. However, as demonstrated above, the ‘‘phase
aqueous phase, only the CQC bonds on the interface could transfer’’ of the catalyst (i.e. from the aqueous phase to the
be hydrogenated. The relatively small amount of the Rh/ polymer phase in this case) is crucial for achieving successful
TPPTS catalyst in the polymer phase during the hydrogenation hydrogenation of large molecular hydrocarbons.

This journal is c The Royal Society of Chemistry 2013 Catal. Sci. Technol., 2013, 3, 2689--2698 2693
View Article Online

Paper Catalysis Science & Technology

3.3 The influence of reaction temperature, pressure,


co-catalyst ligands and co-solvent on NBR latex hydrogenation
The effects of the reaction temperature, pressure and the
co-catalyst ligand on the extent of hydrogenation conversion
were investigated using RhCl(TPPMS)3 and the results are
summarized in Table 3. The hydrogenation reactions were
studied over the range of 85 1C to 130 1C with 0.52 mmol L1
of catalyst loading.
The conversion of NBR is obtained over a reaction period of
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

9 hours at different temperatures. The results clearly showed


that the hydrogenation reaction is sensitive to temperature. The
highest conversion was observed at 100 1C. It was also observed
that the conversion decreased with a further increase in tem-
perature above 100 1C. This is probably due to the occurrence of
catalyst deactivation at high temperature.
This result is further confirmed by considering the NBR Fig. 4 NBR conversions versus reaction time at different temperatures.
Commercial NBR latex = 50 g L1, total volume = 100 mL, RhCl(TPPMS)3 =
hydrogenation profiles shown in Fig. 4, where the highest
0.52 mmol L1, PH2 = 1000 psig.
initial reaction rate was observed at the highest temperature
(130 1C). However, under this reaction temperature a rapid
decay in catalyst activity was observed after 2 hours. The long
catalyst lifetime in the Ruhrchemie/Rhone–Poulenc hydro- the results are presented in Table 3 as well. Compared to the NBR
formylation process below the ceiling temperature of 130 1C hydrogenation in organic solvent, a higher pressure is required in
suggested that the observed rapid deactivation of RhCl- the aqueous latex hydrogenation due to the limited solubility of
(TPPMS)3 in the course of hydrogenation is less likely the result H2 gas in water. As the hydrogen pressure was varied over the
of catalyst thermal decomposition.36 A possibility of the catalyst range of 400 to 1000 psig, the hydrogenation degree increased
deactivation caused by unknown impurities has been taken with increasing hydrogen pressure. However, no significant
into consideration and re-investigated using in-house NBR effect on the catalytic performance of RhCl(TPPMS)3 at high
latex (Section 3.6). hydrogen pressures (>1000 psi) was observed.
The NBR solution hydrogenation system using rhodium It has been found that the co-catalyst ligand PPh3 has a
catalysts such as RhCl(PPh3)3 and RhH(PPh3)4 exhibited a first- significant effect on NBR hydrogenation when using the
to zero-order dependence on hydrogen concentration ([H2]) as the RhCl(PPh3)3 catalyst. Mohammadi et al. pointed out that the
system pressure increased.37 The influence of hydrogen pressure addition of extra free PPh3 maintains the hydrogenation activity
on the present latex hydrogenation reaction was investigated and of RhCl(PPh3)3 by preventing the formation of the non-catalytic
dimeric species [RhCl(PPh3)2]2.38 Wei et al. also found that the
addition of extra PPh3 (e.g. PPh3: RhCl(PPh3)3 = 10 : 1 weight
Table 3 The hydrogenation degree of NBR under different reaction conditions ratio) greatly improved the activity of RhCl(PPh3)3 in NBR latex
hydrogenation as PPh3 promoted the diffusion of RhCl(PPh3)3
[CQCNBR]/
in aqueous latex.11 Therefore, the effect of adding extra TPPMS
Catalyst [Cat.] Temp. Pressure [TPPMS] HD
Entry (mmol L1) (mol/mol) (1C) (psi) (mmol L1) (mol%) to the reaction system was also evaluated in this study. It was
found that the highest catalytic activity was observed without
1 0.52 1100 85 1000 0 74
2 0.52 1100 100 1000 0 95 the addition of any co-catalyst ligand and the activity decreased
3 0.52 1100 115 1000 0 87 after the co-catalyst ligand TPPMS was added, showing that the
4 0.52 1100 130 1000 0 60 NBR latex hydrogenation reaction is negatively influenced by
5 0.52 1100 100 400 0 50
6 0.52 1100 100 600 0 78 increasing the amount of TPPMS. From the sharp decrease of
7 0.52 1100 100 800 0 92 the catalyst activity as a function of increasing TPPMS concen-
8 0.52 1100 100 1200 0 96 tration, the NBR latex hydrogenation may be explained on the
9 0.52 1100 100 1000 1.65 75
10 0.52 1100 100 1000 5.20 62 basis of an equilibrium mechanism.39 The competition for a
11 0.52 1100 100 1000 16.5 49 coordination site on the rhodium (i.e. bi-phosphine rhodium
12 0.52 220 100 1000 0 95
complex) between the added free TPPMS ligand and the CQC
13 0.52 550 100 1000 0 95
14 0.52 1650 100 1000 0 87 bond units of NBR leads to a retardation of the hydrogenation
15 0.52 2200 100 1000 0 73 reaction.40
16a 0.52 1100 100 1000 0 33 Additionally, the influence of nitrile was investigated as the
17b 5.2 1100 100 1000 0 95
polymer loading is varied while holding all other variables
Commercial NBR latex, total volume = 100 mL, reaction time = 9 hours. constant (Table 3, entries 13–15). The dramatic slow down in
a
With the addition of acrylonitrile to the NBR latex before hydrogenation.
The molar ratio of RhCl(TPPMS)3 to acrylonitrile is 1 : 10. b 95% Conver- the hydrogenation reaction with the addition of acrylonitrile to
sion is achieved within 3 hours. the NBR latex (Table 3, entry 16) and the measured low catalyst

2694 Catal. Sci. Technol., 2013, 3, 2689--2698 This journal is c The Royal Society of Chemistry 2013
View Article Online

Catalysis Science & Technology Paper

retention in the polymer phase (35 mol% with addition of the formation of a dihydride.38 However, given that coordina-
acrylonitrile) show that the coordination of nitrile to a catalytic tively unsaturated complexes are often presumed to associate
intermediate existed. with solvent, Joo et al. suggested that the formation of mono-
The distribution of catalysts caused by the interaction hydridorhodium phosphine derivatives [e.g. RhH(TPPMS)3] to
between the catalyst and the concentrated nitrile group in the be a strong possibility [eqn (3)–(5)] which has been extrapolated
NBR polymer chain was also considered. Comparative experi- from the outcome of the hydrogenation of maleic acid and its
ments of hydrogenation of poly-isoprene latex using these two trans-isomer, fumaric acid, in the aqueous phase.44 It has been
catalysts were conducted in our lab and it was found that only known that the addition of an acid or a salt to the polymer latex
RhCl(TPPMS)3 is active for the hydrogenation of poly-isoprene could interrupt its balance and cause polymer precipitation.
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

latex.41 However, no NBR coagulation or precipitation from the latex


was observed using the RhCl(TPPMS)3 catalyst especially under
3.4 Catalytic pathways of the RhCl(TPPMS)3 systems a high catalyst loading (Table 3, entry 17) indicating that little
The catalytic chemistry of Rh(I) phosphine complexes has been HCl or H+/Cl ions were formed during the hydrogenation
extensively studied, resulting in a comprehensive understanding reaction.
of reaction intermediates that bring about the observed reaction
RhCl(TPPMS)3 + H2 " RhH(TPPMS)3 + H + Cl+ (3)
kinetic behavior.42,43 While an extrapolation of this knowledge to
severe reaction conditions may not be straightforward, it is RhCl(TPPMS)3 + H2O " Rh(OH)(TPPMS)3 + H + Cl+ (4)
proposed that a strong correlation exists between the chemistry
Rh(OH)(TPPMS)3 + H2 " RhH(TPPMS)3 + H2O (5)
underlying this work and that documented under mild reaction
conditions. In addition, the coordination of unsaturated complexes with
Fig. 5 illustrates a catalytic mechanism that is consistent solvent was also considered here. Given the observation of the
both with the experimental results shown in the last section transport of the catalyst from the aqueous phase to the polymer
and our understanding of the coordination chemistry of phase during the NBR latex hydrogenation, a consideration that
RhCl(TPPMS)3 in aqueous solution. First, the RhClH2(TPPMS)3 the nitrile group in the polymer chain may inhibit the hydro-
formed by oxidative addition of H2 on RhCl(TPPMS)3 forms the genation cycle is reasonable. As the nitrile group likely coordi-
five-coordinate dihydride according to eqn (2). Then the reaction nates by s donation of its lone pair of electrons, it may
pathway is analogous to the reaction of [RhCl(PPh3)3] in organic therefore compete effectively with olefin for coordination to
solvents which has been well established. unsaturated complexes.
RhCl(TPPMS)3 + H2 " RhClH2(TPPMS)3 (2)
3.5 The effect of adding organic co-solvent
For NBR hydrogenation, the comparative ease of hydrogen The effect of addition of five different organic solvents (mono-
activation compared to addition of a sterically hindered olefin chlorobenzene, toluene, acetone, methyl ethyl ketone and
favors the ‘‘hydride pathway’’ shown in the catalytic mecha- ethanol) was examined and is presented in relation to their
nism. Mohammadi and Rempel showed in their research that effect on the final hydrogenation degree and reaction time of
at 65 1C under 1 bar of H2 the reaction is quantitative towards the respective hydrogenation reactions (Fig. 6). Based on their

Fig. 5 Proposed reaction mechanism for the RhCl(TPPMS)3–NBR system.

This journal is c The Royal Society of Chemistry 2013 Catal. Sci. Technol., 2013, 3, 2689--2698 2695
View Article Online

Paper Catalysis Science & Technology

The fastest hydrogenation reaction was observed with the


addition of ethanol (95 mol% conversion in 3 hours). A similar
phenomenon was observed when alcohol was added for NBR
latex hydrogenation using in situ synthesized RhCl(PPh3)3,31 as
the addition of alcohol increased the efficiency of in situ catalyst
synthesis. Here, the increase in the reaction rate could probably
be explained in that ethanol can counter balance the charges
and reduce the electrical repulsion in the stern-layer of the
micelles which facilitates the diffusion of the catalyst.21 Liu
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

et al. measured the zeta potential of polymer latices containing


various contents of ethanol. They found that the zeta potential
decreased with increased ethanol content showing a reduction
in the electrostatic repulsive energy of the particle surface due
to the presence of ethanol in the dispersing medium.45 This
finding could be useful for the aqueous biphasic reaction
Fig. 6 Influence of organic solvent on the hydrogenation of NBR latex.
especially for improving the catalytic performance within the
Commercial NBR latex with a solid content of 50 g L1, co-solvent/NBR latex = micellar nano-reactor.
1/10 (vol./vol.), total volume = 100 mL, RhCl(TPPMS)3 = 0.52 mmol L1, T =
100 1C, PH2 = 1000 psi. 3.6 Hydrogenation of in-house NBR latex using RhCl(TPPMS)3
Singha et al. first studied NBR latex hydrogenation using
Table 4 The compatibility of different organic co-solvents with NBR and with
the RhCl(TPPMS)3 catalyst. They reported that a maximum of
water 60 mol% conversion could be achieved.23 However, in our
initial investigation, the hydrogenation conversion did not level
Solubility Compatibility off at around 60 mol% but reached more than 95 mol% over a
Organic solvent of NBRa with water
reasonable reaction period. After careful comparison, we found
Mono-chlorobenzene ++  three major differences between our experiments and Singha
Toluene ++ 
Methyl ethyl ketone + + et al.’s results: (1) the type of NBR latex, (2) gel content and (3)
Acetone + + the addition of extra surfactant. As both Singha et al. and
Ethanol  ++ ourselves used commercial NBR latex which normally contains
a
++: good; +: fair; : poor. many additives (e.g. antioxidant, stabilizer), there is a large
possibility that the different hydrogenation behavior could
come from the impact of certain additives in the commercial
compatibility with water and polymers, these five organic NBR latex. We have encountered this issue before when we
solvents could be divided into three categories (Table 4). used different batches of commercial NBR latex. It has been
Mono-chlorobenzene and toluene are good solvents for poly- observed that some of the impurities within the commercial
mers such as NBR but they are immiscible with water. Acetone polymer latex could significantly affect the hydrogenation
and methyl ethyl ketone have certain solubility for NBR and activity of the catalyst or even kill the catalyst.46 Hence, an
compatibility with water. Ethanol is miscible with water but it NBR latex without the addition of additives was synthesized in
cannot dissolve NBR. our lab and used for hydrogenation thereafter. The in-house
First, it is very important to realize that the catalytic system NBR latex was synthesized through emulsion copolymerization
exhibited sufficient catalytic activity (95 mol% conversion in of acrylonitrile and butadiene. A detailed synthesis procedure is
9 hours) without the presence of any additional solvent. Lower summarized in the Experimental Section. The micro-structure
activity of the catalyst was found when mono-chlorobenzene or of our in-house NBR was measured by FT-IR and compared
toluene was used as a co-solvent. With the addition of the with commercial NBR. It was found that the micro-structures
organic solvent, the reaction system changes from a continuous of both NBRs were fairly close. In Fig. 7 the distinct peak
single phase to two phases. The water soluble catalyst stays in exhibited at 2236 cm1 shows the existence of the nitrile
the aqueous phase while the organic solvent swells the NBR group (CRN) in the co-polymer. Due to the existence of two
latex particles to form the organic phase. As a result, the types of CQC bonds, the polymerization of 1,3-butadiene
co-solvent prevents the RhCl(TPPMS)3 from contacting the usually gives rise to three distinct structures, i.e. trans, cis and
CQC bonds in NBR. 1,2(vinyl). The intense peak shown at 969 cm1 belongs to the
Compared to mono-chlorobenzene or toluene, higher con- vibration of the 1,4-trans double bonds, while the small absor-
versions were observed with the addition of acetone and methyl bance at 910 cm1 corresponds to the 1,2-vinyl terminal bonds,
ethyl ketone. This is because the presence of these organic which is much weaker than that of 1,4-trans double bonds.
co-solvents causes the NBR micelles to swell and facilitates the However, the 1,4-cis double bonds which usually show at an
approach of the catalyst to the bulk hydrophobic interior region absorbance around 750 cm1 could not be measured quantita-
of the swollen micelles. tively by FT-IR.

2696 Catal. Sci. Technol., 2013, 3, 2689--2698 This journal is c The Royal Society of Chemistry 2013
View Article Online

Catalysis Science & Technology Paper

around 3, while the molecular weight of highly gelled NBR could


not be directly measured by GPC. The conversions were quite
different for these NBR latices (Fig. 8). Only 45 mol% conversion
could be achieved when using NBR latex containing gel. The
mobility of the polymer chains for these latices is expected to be
quite different. For the gel free polymer, the polymer chains are
mobile under the reaction temperature (TR > Tg). However, the
amount of gel formation caused by cross-linking greatly increases
the molecular weight of the resulting hydrogenated polymer and
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

greatly slows down this movement. It seems that a high gel


fraction had a profound influence on the degree of hydrogenation.
The structure of gelled NBR limits the contact between the catalyst
and the CQC bonds, resulting in low conversion.

4. Conclusions
Fig. 7 Typical FT-IR spectra of NBR from in-house NBR latex. Two types of water soluble rhodium based catalysts, namely,
RhCl(TPPMS)3 and RhCl(TPPTS)3, were studied for NBR latex
hydrogenation. It was found that both catalysts have activity
towards the hydrogenation of NBR latex. The RhCl(TPPMS)3
catalyst was more active than the RhCl(TPPTS)3 catalyst according
to the maximum final conversion in the reaction. The RhCl-
(TPPMS)3 catalyst results in a more than 95 mol% hydrogenation
of the CQC bonds in the NBR latex in the absence of any added
organic solvent. The inherent surfactant propensity of the TPPMS
ligand enables the RhCl(TPPMS)3 complex to diffuse from the
aqueous phase into the NBR phase micelles. The catalyst concen-
tration in the NBR particles during the hydrogenation also con-
firmed that a high degree of hydrogenation can only be achieved
when the catalyst is transported into the solid particles within the
NBR latex. The observed influence of process conditions on the
activity of hydrogenation has led to an extrapolation of a mild
reaction condition mechanism for the RhCl(TPPMS)3 system. In
addition, in-house NBR latex was synthesized and compared with
Fig. 8 Commercial NBR latex and in-house NBR latex hydrogenation by RhCl- commercial NBR latex. High conversion still could be achieved
(TPPMS)3. Commercial and in house NBR latex with a solid content of 50 g L1, using the RhCl(TPPMS)3 catalyst, showing that the RhCl(TPPMS)3
total volume = 100 mL, RhCl(TPPMS)3 = 0.52 mmol L1, PH2 = 1000 psi.
catalyst is probably generically active for NBR latex hydrogenation.
It was also found that the gel fraction in NBR greatly affected the
hydrogenation which could be the reason for the low conversion
The results of the RhCl(TPPMS)3 catalyst for the hydrogena- observed in the previous literature results.
tion reaction of the in-house NBR latex are presented in Fig. 8.
The hydrogenation behavior (hydrogenation curve and final Acknowledgements
conversion) remained practically unchanged for both commer-
cial and in-house synthesized NBR latex at lower temperature, This research work was financially supported by Natural Sciences
demonstrating that the RhCl(TPPMS)3 catalyst is tolerant to and Engineering Research Council of Canada (NSERC), the Canada
additives in the commercial NBR latex under this reaction Foundation for Innovation (CFI) and LANXESS Deutschland
condition. However, the hydrogenation behaviors are signifi- GmbH. Special thanks are also given to University of Waterloo
cantly different when higher temperature was adopted. It was for its support of author’s research.
found that high conversion still could only be achieved in the
in-house NBR latex. References
Second, the effect of gel content of NBR was also investigated.
The commercial NBR latex is free of gel while the NBR latex used 1 M. P. McGrath, E. D. Sall and S. J. Tremont, Chem. Rev.,
by Singha et al. has a gel content of 24 wt%. Therefore, we 1995, 95, 381–398.
synthesized in-house NBR latices with different gel content via 2 G. L. Rempel, Q. Pan and J. Wu, in Handbook of Homogeneous
using different amounts of chain transfer reagent. The in-house Hydrogenation, ed. J. E. C. G. de V. Johannes, Wiley-VCH,
gel free NBR latex has a molecular weight of 75 610 (Mn) with a PDI Germany, 2006, p. 547.

This journal is c The Royal Society of Chemistry 2013 Catal. Sci. Technol., 2013, 3, 2689--2698 2697
View Article Online

Paper Catalysis Science & Technology

3 N. K. Singha, S. Bhattacharjee and S. Sivaram, Rubber Chem. Inorganic Syntheses, ed. M. Y. Darensbourg, John Wiley &
Technol., 1997, 70, 309–367. Sons, Inc., Hoboken, NJ, USA, 1998, pp. 1–8.
4 X. W. Lin, Q. M. Pan and G. L. Rempel, Appl. Catal., A, 2004, 27 W. A. Herrmann, C. W. Kohlpaintner, B. E. Hanson and
276, 123–128. X. Kang, in Inorganic Syntheses, ed. M. Y. Darensbourg, John
5 X. W. Lin, Q. M. Pan and G. L. Rempel, Appl. Catal., A, 2004, Wiley & Sons, Inc., Hoboken, NJ, USA, 1998, pp. 8–25.
263, 27–32. 28 W. Hofmann, Rubber Chem. Technol., 1964, 1–252.
6 G. A. S. Schulz, E. Comin and R. F. de Souza, J. Appl. Polym. 29 D. P. Moraes, M. F. Mesko, P. A. Mello, J. N. G. Paniz,
Sci., 2010, 115, 1390–1394. V. L. Dressler, G. Knapp and É. M. M. Flores, Spectrochim.
7 G. A. S. Schulz, E. Comin and R. F. de Souza, J. Appl. Polym. Acta, Part B, 2007, 62, 1065–1071.
Published on 30 July 2013. Downloaded by University of Saskatchewan on 12/10/2014 14:07:32.

Sci., 2012, 123, 3605–3609. 30 C. B. Ojeda and F. S. Rojas, Talanta, 2007, 71, 1–12.
8 X.-Y. Guo and G. L. Rempel, J. Appl. Polym. Sci., 1997, 65, 31 Y. Liu, Z. Wei, Q. Pan and G. L. Rempel, Appl. Catal., A, 2013,
667–675. 457, 62–68.
9 D. C. Mudalige and G. L. Rempel, J. Mol. Catal. A: Chem., 32 N. Sieffert and G. Wipff, J. Phys. Chem. C, 2008, 112,
1997, 123, 15–20. 6450–6461.
10 D. C. Mudalige and G. L. Rempel, J. Mol. Catal. A: Chem., 33 P. B. Webb and D. J. Cole-Hamilton, in Phosphorus(III)
1997, 116, 309–316. Ligands in Homogeneous Catalysis: Design and Synthesis, ed.
11 Z. Wei, J. Wu, Q. Pan and G. L. Rempel, Macromol. Rapid P. C. J. Kamer and W. N. M. van Leeuwen, John Wiley &
Commun., 2005, 26, 1768–1772. Sons, Chichester, UK, 2012, p. 522.
12 T. Weiss and K. Creutz, US Pat., 20080234437, 2008. 34 V. Kotzabasakis, N. Hadjichristidis and G. Papadogianakis,
13 New HNBR latex for improved heat and oil resistance, J. Mol. Catal. A: Chem., 2009, 304, 95–100.
Sealing Technology, May 2008, p. 2, Retrieved from http:// 35 Y. Dror and J. Manassen, J. Mol. Catal., 1977, 2, 219–222.
www.elsevier.com. 36 J. Herwig and R. Fischer, in Rhodium Catalyzed Hydroformyla-
14 M. S. Spencer and D. A. Dowden, US Pat., 3009969, 1961. tion, ed. P. W. N. M. V. Leeuwen and C. Claver, Rhodium
15 J. Kwiatek, I. L. Mador and J. K. Seyler, J. Am. Chem. Soc., Catalyzed Hydroformylation, 2002, p. 196.
1962, 84, 304–305. 37 J. S. Parent, N. T. McManus and G. L. Rempel, Ind. Eng.
16 F. Joó, Acc. Chem. Res., 2002, 35, 738–745. Chem. Res., 1996, 35, 4417–4423.
17 F. Joó, G. Laurenczy, P. Karády, J. Elek, L. Nádasdi and 38 N. A. Mohammadi and G. L. Rempel, Macromolecules, 1987,
R. Roulet, Appl. Organomet. Chem., 2000, 14, 857–859. 20, 2362–2368.
18 N. Pinault and D. W. Bruce, Coord. Chem. Rev., 2003, 241, 39 F. Joo, L. Somsak and M. T. Beck, J. Mol. Catal., 1984, 24,
1–25. 71–75.
19 E. G. Kuntz, Chem. Technol., 1987, 17, 570–575. 40 A. Bényei, J. N. W. Stafford, Á. Kathó, D. J. Darensbourg and
20 C. W. Kohlpaintner, R. W. Fischer and B. Cornils, Appl. F. Joó, J. Mol. Catal., 1993, 84, 157–163.
Catal., A, 2001, 221, 219–225. 41 P. Piya-areetham, P. Prasassarakich and G. L. Rempel,
21 M. Li, Y. Li, H. Chen, Y.-e. He and X. Li, J. Mol. Catal. A: J. Mol. Catal. A: Chem., 2013, 372, 151–159.
Chem., 2003, 194, 13–17. 42 B. R. James, Homogeneous Hydrogenation, Wiley, New York,
22 D. U. Parmar, S. D. Bhatt, H. C. Bajaj and R. V. Jasra, J. Mol. 1973.
Catal. A: Chem., 2003, 202, 9–15. 43 F. H. Jardine, in Progress in Inorganic Chemistry, ed.
23 N. K. Singha, S. Sivaram and S. S. Talwar, Rubber Chem. S. J. Lippard, John Wiley & Sons, Inc., Hoboken, NJ, USA,
Technol., 1995, 68, 281–286. 1981, pp. 63–202.
24 V. Kotzabasakis, E. Georgopoulou, M. Pitsikalis, 44 F. Joo, P. Csiba and A. Benyei, J. Chem. Soc., Chem. Commun.,
N. Hadjichristidis and G. Papadogianakis, J. Mol. Catal. A: 1993, 1602–1604.
Chem., 2005, 231, 93–101. 45 Z. Liu and H. Xiao, Polymer, 2000, 41, 7023–7031.
25 G. L. Rempel, Q. Pan and J. Wu, US Pat., 7385010, 2008. 46 Z. Wei, Direct catalytic hydrogenation of nitrile butadiene
26 F. Joó, J. Kovács, Á. Kathó, A. C. Bényei, T. Decuir, rubber latex, PhD thesis, Department of Chemical Engineering,
D. J. Darensbourg, A. Miedaner and D. L. Dubois, in University of Waterloo, 2006.

2698 Catal. Sci. Technol., 2013, 3, 2689--2698 This journal is c The Royal Society of Chemistry 2013

You might also like