Wang - Homogeneous Hydrogenation Art of Nitrile Butadiene Rubber A Review, 2013 +
Wang - Homogeneous Hydrogenation Art of Nitrile Butadiene Rubber A Review, 2013 +
Polymer Reviews
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lmsc20
To cite this article: Hui Wang , Lijuan Yang & Garry L. Rempel (2013) Homogeneous
Hydrogenation Art of Nitrile Butadiene Rubber: A Review, Polymer Reviews, 53:2, 192-239, DOI:
10.1080/15583724.2013.776586
Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Polymer Reviews, 53:192–239, 2013
Copyright © Taylor & Francis Group, LLC
ISSN: 1558-3724 print / 1558-3716 online
DOI: 10.1080/15583724.2013.776586
1. General Introduction
In 1930, Konrad and Tshunkur first synthesized acrylonitrile butadiene rubber (NBR)
employing free radical emulsion polymerization.1 Since then, NBR has been widely applied
as an oil-resistant rubber, adhesives, coatings, plastics modifiers and so forth because
of its high strength and excellent elastic properties.2–4 However, due to the presence of
olefinic double bonds, NBR has many drawbacks with respect to chemical, physical and
mechanical properties. The main disadvantage for NBR is its poor aging property, which
results in oxidative and ozonolytic degradation, and the lack of stability to resist extreme
weather conditions. Furthermore, these drawbacks will be aggravated when NBR is used
in manufacturing components of the oil drilling and automotive industries for utilization in
very harsh operation environments. These deficiencies led to the emergence and later on to
the rapid development of hydrogenation techniques for the production of HNBR.
192
Hydrogenation of Nitrile Butadiene Rubber 193
Downloaded by [New York University] at 10:43 17 June 2015
Figure 1. From NBR to HNBR: selective hydrogenation and property improvements (Color figure
available online).
Scheme 1. Catalytic hydrogenation of acrylonitrile butadiene rubber (Color figure available online).
During the last two decades, several comprehensive reviews have provided a thorough
examination on the hydrogenation of elastomers.2–4 These reviews and many other studies
all illustrate that HNBR is a special important elastomer. However, there appears to be no
review particularly concentrating on the advances of the hydrogenation of NBR.
In the past decade a large number of important literature reports have appeared in the
field of hydrogenation of NBR. Many innovative hydrogenation processes were developed,
and perhaps the most striking progress is the development of catalytic hydrogenation of
NBR nanoparticles in the aqueous emulsion/latex media. In addition, a great variety
of robust efficient catalyst systems were designed and applied in the selective reduc-
tion of C C within NBR. Therefore, it is highly desirable to update the knowledge of the
hydrogenation of NBR. The objective of this review was thus to provide an overview of
recent progress towards the homogeneous hydrogenation of NBR.
to be efficient in reducing the gel formation occurring during the diimide hydrogena-
tion process.13 However, the cross-linking (gel formation) still cannot be circumvented
completely.
Catalytic hydrogenation is the most common method for the modification of elastomers,
and its research history can be traced up to seventy years ago.14,15 With respect to substrate
NBR, it has two types of important unsaturated functional groups represented by the olefin
group (C C) and nitrile group (C N) (Scheme 1). Therefore, the central challenge to
produce HNBR is to realize the selective hydrogenation of C C units while keeping
the nitrile group in the rubber intact. Any reduction of the polar nitrile group to imine
or amine would comprise the oil resistant properties for which the end-use application
is desired. Furthermore, such undesirable nitrile hydrogenation is suspected to promote
polymer cross-linking which reduces the processibility of elastomers.
concentration and longer reaction time as well as higher temperature and pressure, which
easily induces the degradation of the polymer and gelation of the reaction system.2
The past decade has witnessed increasing attention concentrated on the homogeneous
catalysts, particularly for their application in the reactions related to polymers. The appli-
cation of the homogeneous catalyst with clearly defined molecular structure allows for a
solid tracking of the catalytic reaction process and interpretation of the mechanism involved
into it, which contributes great favor for the realization of the quantitative hydrogenation,
as well as providing informative inputs for optimizing and controlling of a reaction pro-
cess.16 Homogeneous catalysis is defined as a catalytic system in which the hydrogenation
reaction of the polymer substrate and the catalyst component are brought together in one
phase, in particular the liquid phase. The universal operation of a polymer homogeneous
hydrogenation is similar with a heterogeneous one. First, the polymer, that is, the substrate
is dissolved into a matching organic solvent with abundant solvent power to form a homo-
geneous solution. For example, NBR is preferably dissolved in MCB, acetone, or methyl
Downloaded by [New York University] at 10:43 17 June 2015
ethyl ketone (MEK) while the SBR solution is usually formed in toluene or cyclohexane.
After the substrate polymer solution is prepared, the catalyst in a solid or solution form is
mixed with the polymer solution under a hydrogen atmosphere. Following the completion
of the hydrogenation, subsequent steps include the recycle of the noble metal catalyst such
as Rh if desirable, precipitation, and isolation of the hydrogenated polymer.
Besides its feature of knowability at a molecular level, a homogeneous catalyst has
another two commonly recognized advantages, that is, high selectivity and reactivity. Cat-
alyst reactivity can be enhanced due to the effective dispersion of a homogeneous catalyst
or its derivatives in the substrate solution. Furthermore, homogeneous catalysis offers cat-
alytic mobility and hence circumvents the problem of orienting the long polymer chain on
an immobile active site in the heterogeneous systems.17–19 Although possessing all these
superior features, homogeneous catalytic processes are much less employed in industries
than their counterparts—the heterogeneous ones. The application of a homogeneous cata-
lyst is mostly impeded in the manufacture of products with high economic value. This is
generally blamed to the relatively high production cost of the homogeneous catalyst and
the difficulties encountered for the separation of the catalyst from the final product. The
residue of the expensive catalyst causes an addition of the production cost. Among the
investigated homogeneous catalysts, major concern has been focused on the homogeneous
Rh-based catalysts. Rh complexes representative with Wilkinson’s catalyst are well recog-
nized with their stability and excellent selectivity towards the olefinic C C double bonds in
the presence of the other functionalities. Together with their special properties required for
the unique applications, the recovery of Rh complexes from products is urgently demanded
at the same time. Rh is one of the extremely scarce metals in the earth and its price is
subject to be fluctuated dramatically over the past decade. The historical highest prices
of Rh was recorded as high as 10 000 US $/oz. Therefore, there exists a huge economic
drive for researchers to develop simple and effective methods in order to recover the metal
value after hydrogenation so as to save the high investment costs for purification of the
product with high grade quality, Many substantial improvements have been achieved in the
recovery of noble metal complexes, for example, the use of supercritical carbon dioxide
(scCO2 ) (see Section 5).
In addition to the precious metal based organometallic catalysts, Ziegler catalysts
are another type of homogeneous catalysts, which have been extensively investigated and
applied for catalytic hydrogenation of unsaturated polymers.
Metallocene catalyst involving positively charged Group III or IV metal ions con-
stitutes one type of Ziegler catalyst, which can be used for the polymer hydrogenation.
Hydrogenation of Nitrile Butadiene Rubber 197
Metallocene catalysts are featured by a structure with metal ions sandwiched between
two planar rings of unsaturated carbon atoms that have five atoms apiece. Nevertheless,
metallocenes have a number of disadvantages with respect to the polymer hydrogenation.
They lack thermal stability, cannot be exposed to air, and are intolerant of polar func-
tionality.2 Moreover, the ligands are highly complex and are costly to synthesize. Another
type of Ziegler catalyst is characterized by comprising of a salt and a reducing agent as
effective and necessary constituents. The most used salts for this type of Ziegler catalysts
include organic acid and acetylacetone salts based on the metals of Ni, Co, Pd, Fe, and
Cr, while the two commonly employed co-catalysts are methylaluminoxane (MAO) and
triethylaluminium, (C2 H5 )3 Al.20–21 Compared to the organometallic catalysts comprised
of platinum group metals, Ziegler catalyst systems show the advantages with relatively
inexpensive manufacture cost in terms of the metals used. However, caution should be
taken in that Ziegler catalysts are not applicable for an NBR system attributed to the highly
coordinating nature of the polar nitrile functional group (C N), since the group III cations
Downloaded by [New York University] at 10:43 17 June 2015
are usually extremely Lewis acidic and would be expected to bind irreversibly with the
nitrile functionality.
The history of homogeneous hydrogenation involving organometallic catalysts over
the past half of a century revealed that Rh-based complexes are the most effective and
widely used homogeneous catalysts for the selective hydrogenation of polyolefins, and
are particularly suitable for the commercial production of HNBR. A classic commercial
example is the Tornac R
rubber process, which was invented by Rempel and developed
with Polysar (later Lanxess), for the production of highly saturated acrylonitrile rubbers
using HRhCl(P(C6 H5 )3 )3 and RhCl(P(C6 H5 )3 )3 . The complex is most easily prepared by
treating RhCl3 ·3H2 O with triphenylphosphine (TPP) in ethanol.22–24 Wilkinson’s catalyst
has numerous advantages compared to other catalyst systems: (i) high selectivity toward
the olefin double bond; (ii) effective suppression of crosslinking problems; (iii) easy to
achieve high conversion and (iv) excellent air stability in the solid form.2
Besides the studies focusing on the recovery of Rh catalysts, a lot of effort has been
made by researchers to design and develop novel homogeneous catalysts based on the less
expensive metals, targeting at seeking proper catalysts for manufacture of high performance
HNBR with a reduced cost. These mainly include Ru-, Os-, Ir-, and Pd-based catalyst
systems. Nevertheless, some drawbacks were found when these catalysts are applied for
the homogeneous hydrogenation reactions (as listed in Table 1), which thus hinders the
wide application of these metal catalysts in industry.
Table 1
The drawbacks of noble metal based catalysts compared with Rh complexes in the homo-
geneous hydrogenation of olefinic unsaturated polymer
the present time, the solution hydrogenation process of HNBR is marketed by two major
companies in the world. Lanxess (formerly Polysar/Bayer) produces HNBR via a homoge-
neous hydrogenation process in Texas (US) and Leverkusen (Germany) while Nippon Zeon
manufactures HNBR using a heterogeneous hydrogenation process in Japan. However, the
solution hydrogenation process is known to bear several disadvantages, and the major con-
cern that has attracted considerable attention involves the usage of large amounts of organic
solvents which not only increases the cost of the process but also raises environmental
concerns.
In most cases for the solution hydrogenation of unsaturated polymers, in no way limited
to NBR, a fairly universal procedure is implemented (Fig. 2).2 First, the substrate polymer
is dissolved in a suitable solvent with a low concentration of ranging 2–10 wt% in order
to reduce the viscosity of polymer solution. MCB is a practical solvent and MEK is also
effective as the solvent under the mild reaction conditions such as ambient pressure and low
temperature (<100◦ C). After the polymer solution is ready, the hydrogen or an inert gas like
nitrogen was used to purge the reactor to remove air. The reaction system is subsequently
heated up to the desired temperature (preferred > 100◦ C) and then the hydrogen is charged
to reach a certain pressure (preferred > 2.7 MPa). Depending on the specific condition, the
catalyst can be purged into the reaction media before or after reaching the set up temperature
and pressure with or without addition of co-catalyst like TPP. Following the completion
of the hydrogenation, the catalyst and solvent are separated, recovered, and recycled. The
hydrogenated polymer is finally precipitated, dried, and baled.
Mohammadi and Rempel established the relationship between the quantitative hy-
drogen consumption and the reaction time without the influence of disturbance which
is inherent in normal sampling operations by using of a computer-controlled gas-uptake
equipment.28 The reaction experiments were investigated in three types of organic solvent
including MEK, MCB, and ethyl acetate (EtOAc) using Wilkinson’s catalyst under mild
conditions with moderate temperatures of over the range of 20–50◦ C and hydrogen pres-
sures below 1 atmosphere. Their results indicated that the selectivity of Wilkinson’s catalyst
for internal double bonds and terminal double bonds was greatly influenced by the nature
of the organic solvent used. In MEK and EtOAc, the catalyst showed nonselectivity for
the internal and terminal double bond. However, in the experiments of using MCB as the
Hydrogenation of Nitrile Butadiene Rubber 199
Downloaded by [New York University] at 10:43 17 June 2015
Figure 2. Schematic illustration of industrial process (i.e., solution hydrogenation) for the production
of HNBR.2 Reproduced from Rempel et al.2 with permission from Wiley.
solvent, the catalyst exhibited the distinct preference towards hydrogenation of the vinyl
groups. It is known that the polymerization of butadiene (BD) can yield three types of
C C, namely, trans, cis, and vinyl. The concentration and distribution of different types of
C C in the latex particles has a definite influence on the hydrogenation rate and reaction
kinetics since they have different reactivities. Generally, the order of the reactivity of these
three types of C C is vinyl > cis > trans using an Rh catalyst29 while the order changes to
vinyl > cis ≈ trans in the hydrazine/oxidant-induced diimide hydrogenation reaction.30–33
The vinyl-C C are more reactive towards chemical modification reactions than 1, 4 units
due to their lower steric hindrance.
More importantly, the authors found that the hydrogenation rate exhibited an apparent
first-order in the [Rh], [C C] of NBR, and [H2 ] at low hydrogen pressure, as expressed in
Eq. (1).34 Apparent activation energy (Ea ) of 87.28 kJ/mol was thus obtained.
k KK K1 [H2 ][Rh][C=C]
RH = × 100%
KK1 + K [PPh3 ] + KK [H2 ][PPh3 ] + KK K1 [H2 ] + KK1 K5 [C≡N] + KK1 K2 K [H2 ][C≡N]
(1)
where RH is the hydrogenation reaction rate; k is the kinetic rate constant, and
K,K ,K1 ,K2 ,K5 are the equilibrium constants.
In addition, an inverse dependence of reaction rate on the nitrile concentration [C N]
was observed. Based on these kinetic results and experimental observations a reaction
mechanism describing the hydrogenation of NBR using RhCl(PPh3 )3 was put forward as
shown in Scheme 2.28,34,35 It is worth noting that Scheme 2 can also be used to describe the
kinetics and mechanism of the hydrogenation reactions of BR and SBR, where K 2 and K 5
200 H. Wang et al.
Downloaded by [New York University] at 10:43 17 June 2015
vanish. The kinetic constants and equilibrium parameters of several typical polymer/solvent
systems are presented in Table 2.
Bhattacharjee et al. extended Mohammadi and Rempel’s work28 to a higher reaction
temperature of 100◦ C and pressures up to 5.6 MPa in an effort to optimize the process and
characterize the final product.38 The method reported to obtain the data involved sampling
of the polymer solutions during the experimental runs at various time intervals, followed by
IR, NMR, and iodometric analyses to determine the degree of hydrogenation. The effect of
various reaction factors such as temperature, time, hydrogen pressure, catalyst concentra-
tion, and solvent were comprehensively examined. Under optimum reaction conditions a
100% hydrogenation degree was achieved at a catalyst concentration of 0.02 mmol, under
5.6 MPa hydrogen pressure at 100◦ C in MCB for 11 h. It was observed that the tempera-
ture plays a crucial role in the hydrogenation reaction. There was no conversion at room
temperature. However, the degree of hydrogenation increased with an increase in tempera-
ture, hydrogen pressure and RhCl(PPh3 )3 concentration. They showed that MCB is a better
solvent for the NBR solution hydrogenation using Wilkinson’s catalyst compared with
ODCB and monobromobenzene. Through the kinetic studies of the NBR hydrogenation,
the activation energy was calculated as 22 kJ/mol, dramatically smaller that the value of
87.3 kJ/mol reported by Mohammadi and Rempel28 and also not consistent with the normal
range of reaction activation energy as shown in Table 2. It probably was a result that the
hydrogenation reaction would became diffusion-controlled under the reaction conditions
employed in Bhattacharjee’s study.38 Some important physical properties of HNBR like
molecular weight, glass transition temperature (T g ) and stress-strain behavior were also
reported.
Downloaded by [New York University] at 10:43 17 June 2015
Table 2
Kinetic constants and equilibrium parameters of several typical polymer/solvent systems.
Polymer Solvent K [mM−1] K 1 [mM] K’ [mM−1] K 2 [mM−1] K 5 [mM−1] k’ [mM−1·s−1] Ea [kJ/mol] References
NBR MCB ∞ 1.44 3.41 × 10−3 3.98 × 10−2 2.71 × 10−2 1.19 73.5 24
NBR
MEK ∞ 0.198 0.276 6.5 × 10−2 0 4.23 × 10−4 87.3 28
1,4-PB 1,2-Dichlorobenzene 0.60 4.45 0.59 0 0 1.28 × 10−4 98.5 36
(ODCB)
SBS ODCB 0.31 3.13 0.63 0 0 3.26 × 10−4 78.8 37
SBR Toluene 1.23 4.70 0.72 0 0 4.77 × 10−4 60.8 37
201
202 H. Wang et al.
Table 3
Hydrogenation of NBR with cationic rhodium complexes.39
Mao and Rempel reported a class of cationic rhodium catalysts with the form of
[Rh(diene)(L2 )]+ as shown in Table 3, where diene is norbornadiene (NBD) or 1,5-
cyclooctadiene (COD) and L is TPP or L2 is a chelating phosphine ligand, for the hy-
drogenation of NBR under 3.42 MPa hydrogen pressure at 140◦ C in MCB.39 As indicated
in Table 3, among the different types of ligands used, the activity of [Rh(diene)(PPh3 )2 ]+
is similar to RhCl (PPh3 )3 (Wilkinson’s catalyst), which showed that the catalytic activity
of the [Rh(diene)(L2 )]+ complexes is mainly dependent on the nature of phosphine. The
catalytic activity and a detailed kinetic study of this series of complexes were carried out
by screening groups of experiments. The reaction was monitored by periodically with-
drawing samples from the reactor and recording their IR spectra to determine the extent of
hydrogenation.
As shown in Table 3, Rh(NBD) (PPh3 )2 + showed a very high catalytic activity and
a near completion of hydrogenation. A mechanism was proposed to describe the kinetics
of hydrogenation of NBR using Rh(NBD) (PPh3 )2 + as the catalyst (Scheme 3). An Ea
calculated as 60.7 kJ/mol was acquired, which shows a satisfactory comparability with
the other values of Ea found for RhCl (PPh3 )3 (73.5 kJ/mol in MCB24 and 87.3 kJ/mol
in MEK28) and for RhH (PPh3 )4 (57.4 kJ/mol in MCB24). These data indicate that the
hydrogenation rate of these systems is controlled by the chemical reaction, rather than the
mass diffusion of the catalyst, polymer, or hydrogen, because the diffusion control reaction
generally requires Ea < 18 kJ/mol.
The role that TPP played in the hydrogenation of NBR using RhCl(TPP)3 (that is,
RhCl(PPh3 )3 ) in MCB was carefully examined by McManus and Rempel.40 The study
indicated that a side reaction occurred in parallel with the hydrogenation, which produces
benzene and HCl from the solvent of MCB. The by-products were likely generated by
the oxidative addition of the C Cl bond in MCB to a Rh intermediate complex during
the catalyst cycle for the hydrogenation (Scheme 2 of mechanism of hydrogenation using
RhCl(PPh3 )3 ), followed by reductive elimination of benzene in conjunction with hydrogen
addition to the Rh center. This reaction will result in the formation of a less active Rh
Hydrogenation of Nitrile Butadiene Rubber 203
Downloaded by [New York University] at 10:43 17 June 2015
intermediate, which would induce rapid deterioration in catalyst activity. They reported
that excess free TPP could not only prevent a degradative side reaction of the active Rh
center, but act as a base to neutralize the HCl generated in the process. The relationship
between the addition of TPP and the hydrogenation degree is shown in Table 4. Parent
et al. also reported that addition of excess free TPP to the hydrogenation system under
elevated temperature and pressure could stabilize the Rh complex and prevent reduction of
RhCl(TPP)3 to Rh metal.24
McManus and Rempel proposed a mechanism to account for the formation of benzene
and HCl.40 The concentration of C C is considered to be the most important factor to affect
the amount of benzene produced during the hydrogenation process. Due to the neutralization
Table 4
The effect of TPP concentration on the degree of hydrogenation of NBR
using RhCl (TPP)3 .40
reaction of TPP and HCl, alternate bases used as co-catalyst was added into the reaction
system to consume the HCl, by which the TPP consumption could be reduced and the overall
catalyst activities compared to the chosen concentration of RhCl(PPh3 )3 were improved. The
results showed that amines, especially primary amines and metal oxides such as MgO and
CaO, along with low levels of TPP could afford effective hydrogenation; however, these
co-catalysts are not effective without the presence of TPP.
Hsu et al. examined a bimetallic complex catalyst system (Rh-Ru) for NBR hydro-
genation and this bimetallic catalyst is designed to be comparable to the conventional single
Rh catalysts in activity and selectivity.41 Thus, the cost of Rh would be offset to a certain
extent by the relatively lower price of Ru. The bimetallic complex catalyst system is pre-
pared based on a combination of Rh and Ru with different mole ratios from 1:1 to 4:1. In
addition, the reaction is conducted by using a two-stage temperature reaction procedure. A
degree of hydrogenation of 96.3% was achieved when the amount of Rh and Ru is equal
at the two-stage temperature of 110◦ C for 2.5 h and 145◦ C for 1.5 h under 0.8 MPa H2
Downloaded by [New York University] at 10:43 17 June 2015
pressure. When the ratio of Ru/Rh was increased to 4:1 (mol ratio), 99% hydrogenation
degree was obtained under the same experimental conditions. If only RuCl2 (PPh3 )3 is em-
ployed for the hydrogenation, gel formation is observed, which is not found when using
the bimetallic complex catalyst system (wherein Rh/Ru ≥ 1). The increase in the catalytic
activity of bimetallic catalyst system may be ascribed to a synergistic effect between two
metal complexes.
Scheme 4. Schematic diagram illustrating the preparation and catalytic latex hydrogenation of diene-
based polymer nanoparticles. The latex hydrogenation using a catalyst pre-dispersion approach, which
consisted of three steps: the addition of catalyst (a-b), introduction of hydrogen gas (b-c), and selective
reduction of olefinic groups in nanoparticles (c-d).51 Reproduced from Wang et al.51 with permission
from Wiley (Color figure available online).
process for the parent polymer synthesis, as the conventional route for synthesis of diene-
based polymers is a semibatch emulsion polymerization process and its convenient products
are in latex form. The implementation of latex hydrogenation is not uniquely confined to
NBR and most unsaturated rubber latex can be hydrogenated directly in the latex form. On
the other hand, the recovery/recycle operation of the catalyst from the resultant hydrogena-
tion media will benefit greatly from the utilization of the latex hydrogenation,43–44 which
will be introduced in Section 3.2.3. Therefore, the commercialization of latex hydrogena-
tion is currently standing as the pursuit of rubber industry partners and representing an
important active research topic in academic community.3,4, 45–50
The uniqueness of the latex reduction process lies in that each latex particle functions
as an ideal reactor and the hydrogenation of NBR occurs inside each suspended particle
(Scheme 4).51 Latex hydrogenation could be classified into two categories according to the
nature of catalyst employed, namely an oil soluble catalyst and a water soluble catalyst. In
some cases, a suitable organic solvent is required to assist in dissolving the catalyst and
subsequently dispersing the organic solvent system in water.
3.2.1 Utilization of Oil Soluble Catalysts. The first example of NBR emulsion hy-
drogenation was reported by Weinstein in 1984 using Wilkinson’s catalyst. However,
the product was not HNBR, but another elastomer of being acrylonitrile-ethylethylene-
tetramethylene copolymer. The hydrogenation was carried out at 70–75◦ C and an initial
206 H. Wang et al.
Figure 3. The TEM of commercial NBR latex nanoparticles. This NBR latex has a solid content
Downloaded by [New York University] at 10:43 17 June 2015
of ∼ 12 wt% and a mean diameter of about 70 nm (L); latex hydrogenation of commercial latex
particles. The hydrogenation temperature is set up as 145◦ C and hydrogen pressure is maintained
under 6.8 MPa (R).56 Reproduced from Wei et al.56 with permission from Wiley.
pressure of 240–275 kPa.52 Mahittikul et al. hydrogenated natural rubber latex (NRL)
using OsHCl(CO)(O2 )(PCy3 )2 53 or [Ir(cod)(PCy3 )(py)]PF6 54 as catalyst, respectively. A
higher conversion of greater than 95 mol% was achieved in both catalyst systems under
proper experimental conditions. The MCB was required in these two studies in order to
reach a desired degree of hydrogenation. Chemtob et al. studied the direct hydrogenation
of purely linear polybuta-1,4-diene (1,4-PB) latex in the presence of RuCl2 (PPh3 )3 and
complete hydrogenation was observed in 8 h using a catalyst loading around 10 wt% over
polymer substrate under 20 MPa hydrogen at 130◦ C.55 Rempel and coworkers developed
a NBR latex hydrogenation technique by which a NBR latex was directly hydrogenated
without requiring pretreatment before hydrogenation and no organic solvent was used.56
Wilkinson’s catalyst and TPP were used as a catalyst and co-catalyst ligand respectively.
In this study, the added excess TPP plays a crucial role to transport the catalyst into the
latex particles; thereby, no alien solvent was used. The hydrogenation of latex was carried
out at different pressures, temperatures and catalyst concentrations. More than 95 mol%
hydrogenation degree, without any cross-linking problem could be achieved at 145◦ C and
6.8 MPa. It should be noted that 95% conversion without gel formation is an important
criteria for the commercialized hydrogenation of NBR, because it is a critical level that
guarantees the outperformance of HNBR in miscellaneous applications and facilitates the
successive processibility of HNBR, for example, vulcanization. During the hydrogenation
reactions, the degree of hydrogenation increases with increasing catalyst concentration and
reaction time (Fig. 3). No coagulation was observed even under the running conditions of
longer reaction time of above 90 h. It is interesting to find that the particle diameter and
its distributions of original NBR latex particles are maintained consistent with HNBR after
hydrogenation. The major drawback of this study is that the hydrogenation rate is quite
slow (Fig. 3), which became the main obstacle limiting the further development of this
technique.
Very recently, Lanxess (Sarnia, Canada) filed a patent focusing on the latex hydro-
genation of NBR latex using in-situ synthesized RhCl(PPh3 )3 .57 The typical experimental
procedure was reported as: first mix 25 mL latex, 75 mL water, and a certain amount of
Hydrogenation of Nitrile Butadiene Rubber 207
Downloaded by [New York University] at 10:43 17 June 2015
Figure 4. Size and morphology of NBR nanoparticles observed by TEM imaging. (a) NBR nanopar-
ticles; (b) the hydrogenated NBR nanoparticles (HNBR) with the hydrogenation degree higher than
95 mol%. Figures 4a and 4b represent the same batch nanoparticles before and after hydrogenation.
Experimental conditions of synthesis: AN = 5 mL, BD = 15 mL, APS = 0.2 g, distilled water =
80 mL, T = 50◦ C, addition time = 200 min, aging time = 8 h, concentration of GS 12–3-12 is
50 mM. Experimental conditions of latex hydrogenation: RhCl(PPh3 )3 /NBR is 1 wt%, RhCl(PPh3 )3
/TPP is 10 wt%, 600 rpm of agitation, 130◦ C, 6.89 MPa of H2 .51 Reproduced from Wang et al.51 with
permission from Wiley (Color figure available online).
RhCl3 ·3H2 O and TPP. The mixture was then placed inside a 300 mL glass lined stainless
steel autoclave, degassed with nitrogen, and heated up to a setup temperature. The hydro-
genation embarked on when the hydrogen was charged. The RhCl3 ·3H2 O was pre-mixed
with TPP before contacting with NBR latex as well as one experiment without addition of
TPP was carried out. The results showed that when RhCl3 ·3H2 O was used in the absence
of TPP, gel was generated and the hydrogenation degree can only reach 23 mol% at 100◦ C
and 6.8 MPa hydrogen pressure after 4 h. In case of the presence of TPP, the hydrogenation
situation was greatly improved. 97 mol% degree of hydrogenation could be obtained at
160◦ C and 6.8 MPa with a TPP/RhCl3 ·3H2 O ratio of about 27 after 19 h and no gel was
observed. The catalyst was synthesized in-situ; therefore, no alien organic solvent was used
to transport the catalyst.
These studies mentioned above confirm that the central challenge inherent in the
field of latex hydrogenation is the optimal interplay of accelerating the hydrogenation rate,
decreasing the required quantity of catalyst, and eliminating the need for an organic solvent,
which is the key for success in this latex hydrogenation route. Rempel and coworkers have
met this challenge completely through integrating the advantages of polymer nanomaterials.
They found that the latex hydrogenation rate increased drastically with decreasing of the
particle size of the substrate particles.51 For the first time, they synthesized a type of NBR
nanoparticles with a size of less than 20 nm with narrow particle size distributions via a
semibatch microemulsion polymerization method using Gemini surfactant trimethylene-
1,3-bis (dodecyldimethylammonium bromide) (GS 12-3-12) as the surfactant (Fig. 4).
Through programming the concentration of surfactant, the particle size of NBR particles
can be controlled in the range from 17–54 nm. The prepared NBR nanoparticles were then
used as the substrates for the catalytic latex hydrogenation catalyzed by RhCl(PPh3 )3 . For
208 H. Wang et al.
Downloaded by [New York University] at 10:43 17 June 2015
Figure 5. The dependence of latex hydrogenation rate upon the particle size. Experimental conditions
of preparation: AN = 5 mL, BD = 15 mL, APS = 0.2 g, distilled water = 80 mL, T = 50◦ C, addition
time = 200 min, aging time = 8 h, concentration of surfactant is a variable. Experimental conditions of
latex hydrogenation: RhCl(PPh3 )3 /NBR is 1 wt%, RhCl(PPh3 )3 /TPP is 10 wt%, 600 rpm of agitation,
130◦ C, 6.89 MPa of H2 .51 Reproduced from Wang et al.51 with permission from Wiley (Color figure
available online).
the 17 nm NBR nanoparticles, the hydrogenation rate is maximized and a near completion
of hydrogenation was reached within 18 h using only 0.1 wt% RhCl(PPh3 )3 based on the
mass of NBR solid (Figs. 5 and 6). It is worth pointing out that no organic solvent was used
in the latex hydrogenation process and no gel formation was observed.
Furthermore, Rempel and coworkers prepared a type of PMMA-NBR core-shell struc-
tured nanoparticles using a two-stage semibatch microemulsion polymerization method
using GS 12-3-12 as the surfactant (Fig. 7).58 As shown in Fig. 7 the light domains corre-
spond to the NBR shell whereas the PMMA core exhibits dark regions. The size of PMMA
is 22.7 and PMMA-NBR is 30.6 nm (DLS data). In the second stage of their research, the
prepared PMMA-NBR nanoparticles were hydrogenated in latex form using RhCl(PPh3 )3
in the absence of organic solvent. It was found that the hydrogenation rate is very fast.
Without using organic solvent, the hydrogenation of NBR shell was near completed using
3 h with 0.9 wt% RhCl(PPh3 )3 at 130◦ C and hydrogen pressure of 6.89 MPa. Because the
NBR layer in the core-shell nanoparticles is very thin with a thickness of around 5 nm,
which provides a considerable benefit for the mass diffusion of the catalyst into the NBR
polymer layer. In contrast, the commercial NBR nanoparticles with a larger particle size
(∼ 70 nm) (Fig. 3) showed a very slow rate of reaction as a result of severe mass transfer
limitation of catalyst inside the particles.
The synthesis and following green hydrogenation process of NBR nanoparticles and
PMMA-NBR core-shell nanoparticles provided by Rempel and coworkers can be extended
to the other latices made from semibatch microemulsion polymerizations containing other
unsaturated polymers such as SBR and BR. These two studies confirm the significance in
Hydrogenation of Nitrile Butadiene Rubber 209
Downloaded by [New York University] at 10:43 17 June 2015
Figure 6. Effect of Wilkinson’s catalyst amount on the latex hydrogenation rate. Experimental
conditions of preparation: AN = 5 mL, BD = 15 mL, APS = 0.2 g, distilled water = 80 mL, T =
50◦ C, addition time = 200 min, aging time = 8 h, concentration of surfactant is 50 mM. Experimental
conditions of latex hydrogenation: RhCl(PPh3 )3 /TPP is 10 wt%, 600 rpm of agitation, 130◦ C, 6.89
MPa of H2 .51 Reproduced from Wang et al.51 with permission from Wiley (Color figure available
online).
Figure 7. TEM imaging and 1H NMR spectra of PMMA-NBR and PMMA-HNBR core-shell
nanoparticles. Hydrogenation conditions: RhCl(PPh3 )3 /NBR shell is 0.9 wt%, RhCl(PPh3 )3 /TPP
is 10 wt%, 600 rpm of agitation, 130◦ C, 6.89 MPa of H2 , reaction time = 3 h. Reproduced from
Wang et al.58 with permission from Wiley (Color figure available online).
210 H. Wang et al.
designing and synthesizing the nanoscale polymeric materials and show great promise in
the large scale production of green hydrogenation of NBR in the latex form.
3.2.2. TPP’s Role as a Catalyst Carrier. The unique role of TPP in the typical solution
hydrogenation process pertains to the maintenance of the superior catalytic efficiency of
RhCl(PPh3 )3 . With respect to the latex hydrogenation, the free ligand TPP bears an ex-
tended important role: being a carrier to deliver the oil miscible RhCl(PPh3 )3 molecules
from the aqueous regime to the polymer regime (particles). This phase-transfer process was
described in Scheme 5. In a typical run for catalytic latex hydrogenation, the RhCl(PPh3 )3
and TPP with a ratio of 10:1 (w/w) were first added into the latex (Scheme 5a) before
charging H2 gas. Under elevated reaction temperatures (generally above 90 ◦ C), TPP (m.p.
∼80 ◦ C) will be melted and then form numerous liquid droplets, which will act as a con-
tinuous phase to dissolve the catalyst to form a homogeneous mixture solution (Scheme
5b). This mixture will be stabilized and suspended in the latex system by the surfactant
Downloaded by [New York University] at 10:43 17 June 2015
molecules (Scheme 5c). In a latex system, the surfactants are moving in a dynamic assem-
bling state and the surfactant molecules are actively being assembled onto the particles
and departed from them. These surfactant molecules will bring the TPP-catalyst mixture
droplets into the latex particles (Scheme 5d). The liquid state TPP is miscible with the
polymers like NBR and HNBR. Therefore, when RhCl(PPh3 )3 and TPP are transferred
into polymer phase, the compatibility between the catalyst and polymer phase is promoted
by TPP and finally, RhCl(PPh3 )3 is dispersed inside the polymer matrix. Form the view-
point of this mass transfer mechanism, these three components TPP, catalyst, and polymer
substrate are considered to form a homogeneous mixture under the proper experiential
conditions. It can be stated that the latex hydrogenation described here can be regarded as
“mini-bulk” hydrogenation, i.e., each latex nanoparticle is the location for the solid bulk
hydrogenation.
3.2.3. Utilization of Water Soluble Catalysts. The use of hydrophilic analogues of Wilkin-
son’s catalyst have been attracting attention over the past few years in the area of the
biphasic hydrogenation of olefinic unsaturated polymers since it provides a chance to
separate the end-product from the catalyst in the homogeneous catalytic systems. In
addition, water is used as the bulk reaction solvent instead of organic solvents, which
thereby offers a safer, cheaper, and environmentally benign alternative for current industrial
processes.59–61
Kuntz introduced the idea of biphasic catalysis in water and the first industrial appli-
cation of biphasic hydroformylation promoted the interest in such water-soluble organo-
metallic systems.62 Nomura has summarized transition metal catalyzed hydrogenation in
water.63 Amongst, the most important and efficient water-soluble catalysts is RhCl(TPPTS)3
(TPPTS = trisulphonated-triphenylphosphine, P(C6 H4 -m-SO3 Na)3 ).64 Following this suc-
cess, many water-soluble ligands and catalysts were studied for the biphasic reaction with
a polymeric substrate.
The first water soluble catalyst that applied for catalytic hydrogenation of NBR la-
tex is RhCl(DPM)3 65 (DPM = diphenyl phosphino benzene m-sulphonat, prepared by
Kalck and Montail66), which was synthesized and examined on its performance by Sivaram
and coworkers. A large range of operational conditions including pressure, temperature,
and catalyst concentration were investigated with respect to their influence on the cat-
alytic hydrogenation process of the NBR latex via RhCl(DPM)3 . More than 60% con-
version could be achieved with 0.5 mol% catalyst in 12 h at 75◦ C under atmospheric
Hydrogenation of Nitrile Butadiene Rubber 211
Downloaded by [New York University] at 10:43 17 June 2015
Scheme 5. Mass transfer of catalyst using TPP as the catalyst carrier from aqueous phase to bulk
polymer phase (Color figure available online).
H2 pressure. The latex was found to remain stable during the process of catalytic reac-
tion, accompanied by an increase in gel content but no coagulation of the latex. Accord-
ingly, the average diameter of the NBR particle in the latex was maintained at a simi-
lar level throughout the hydrogenation, but the particle size distribution showed a slight
broadening compared with the original latex. Elevating temperature and H2 pressure was
found to be effective for increasing the conversion of C C at no cost of increasing gel
content.
The catalytic reaction rate in such systems is largely determined by the phase contacting
rate of the catalyst and the substrate molecules which are soluble in separate phases and are
scarcely miscible with each other. Therefore, the catalytic reaction rate becomes extremely
slow due to the phase transfer limitations.61
212 H. Wang et al.
In order to speed up the phase transfer rate of the catalyst or the polymer molecules
through the W/O interface, several attempts were made to meet this goal. Mudalige et al.
examined the catalytic activity of two water soluble Rh complexes for hydrogenation
of various polymers involving BR, SBR, and NBR in an aqueous/organic biphasic sys-
tem at 100◦ C and 5.5 MPa H2 .61 Instead of the usage of the polymer latexes as re-
action media, polymer emulsions were prepared by suspending polymer solutions (in
toluene) directly into distilled water and applied as a W/O biphasic system. The re-
sults were compared between two catalysts named [RhCl(HEXNa)2 ]2 (HEXNa = Ph2 P-
(CH2 )5 CO2 Na) and RhCl(TPPMS)3 (TPPMS = monosulphonated-triphenylphosphine,
PPh2 (C6 H4 -m-SO3 Na)). Both complexes showed a reasonable catalyst activity toward poly-
mer hydrogenation in a W/O biphasic media. The hydrogenation conversion of NBR latex
obtained after 6 h was 53% at 100◦ C and 5.5MPa. Nevertheless, the catalyst was found to
leach into the organic phase with proceeding of the reaction.
Yasuhiko et al. used another water soluble complex Rh-TPPTS for catalytic hydro-
genation of NBR latex.67 The produced latex after the hydrogenation reaction was stable
Downloaded by [New York University] at 10:43 17 June 2015
and the hydrogenation conversion was 56% after 24 h at 60◦ C and under 0.8 MPa H2 .
Lanxess (Sarnia, Canada) disclosed a patent concentrating on using the water-soluble
catalyst RhCl(TPPMS)3 for hydrogenation of NBR latex with addition of a water-insoluble
co-catalyst TPP in the absence of any organic solvent.68 The experimental procedures are
similar with those previously described in another patent claimed by Lanxess except for
the replacement of RhCl3 ·3H2 O by RhCl(TPPMS)3 .57 The addition of TPP was found
to be important as for suppressing the cross-linking and improving the final degree of
hydrogenation. The results showed that only 43 mol% degree of hydrogenation was obtained
after 21 h at 160◦ C under 6.8 MPa H2 and gel was generated without the presence of TPP.
However, after the addition of TPP (around ten times over catalyst in mass), a 92 mol%
degree of hydrogenation was achieved under the same experimental conditions and no gel
was generated under the same experimental conditions.
Wei et al. studied another biphasic catalytic system comprised of a polyether modified
ammonium salt ionic liquids and organic solvent as reaction media regarding its applica-
tion of hydrogenation of block copolymers SBS using Ru/TPPTS complex.69 They found
that the addition of TPP worked effectively as for the particular improvement of the hy-
drogenation degree of SBS. A high hydrogenation degree of 89 mol% was reported for
the polybutadiene segments of SBS under the optimized conditions of Ru/TPPTS/TPP
= 1: 5: 2 at 150◦ C under 5.0 MPa H2 . The immobilization of the Ru catalyst in the
ionic liquid greatly promoted the recycle of the catalyst by phase separation. However,
more work should be conducted towards enhancing the hydrogenation degree, increas-
ing the hydrogenation rate, and improving catalyst-recycling efficiency. The catalyst of
RhCl(TPPTS)3 was also investigated for its efficiency in catalytic hydrogenation of several
polymers involving BR, NBR and SBR in a ionic liquid biphasic system of toluene/N,N -
butylmethylimidazolium tetrafluoroborate system at 100◦ C and 3.1 MPa.70 The activity
of the catalyst in the ionic liquid differs among polymer substrates. BR can be some-
how hydrogenated directly via the catalyst, whereas co-solvent water has to be added for
enhancing the solubility of the catalyst in the ionic liquid for the substrate of NBR and
SBR.
Very recently, Papadogianakis and coworkers studied the latex hydrogenation of BR
using a water-soluble complex RhCl(TPPTS)3 .71 The authors reported that RhCl(TPPTS)3
is very active for the BR hydrogenation under the studied mild reaction conditions. The
turnover frequency (TOF) can reach 1245 h−1 through adjusting the experimental conditions
Hydrogenation of Nitrile Butadiene Rubber 213
and a conversion of 83 mol% was obtained at this TOF. Different types of surfactants and
surfactant combinations were used to emulsify the n-hexane-PB solution for the purpose
of investigating the effects of added surfactant on the catalytic activity of RhCl(TPPTS)3 .
These emulsified systems include single cationic surfactant dodecyltrimethylammonium
chloride (DTAC) or cetyltrimethylammonium chloride (CTAC), combination of non-ionic
Brij type surfactants with DTAC or CTAC, and mixed surfactants comprising of anionic
surfactant sodium dodecylsulfate (SDS) and DTAC. It was found that the types of surfactant
imposed a significant effect on the activity of the catalyst and thereby the hydrogenation
rate. The authors attributed the influence of the nature of different surfactant systems on
the catalytic activity of Rh/TPPTS to the pronounced effect of the hydrophilic-lipophilic
balance (HLB value) on the distribution of catalytically active Rh of RhCl(TPPTS)3 inside
the latex particles. They proposed a schematic model to illustrate the location of catalyst
in the latex particles, as shown in Scheme 6. Therefore, it is not hard to understand
that the pH of the reaction media should induce a sound effect on the activity of the
Downloaded by [New York University] at 10:43 17 June 2015
catalyst. It was reported that high TOFs were achieved under neutral conditions whereas
the acidic or basic conditions would render negative effect on the catalytic activity. In
another report of Papadogianakis and coworkers, they examined the latex hydrogenation of
diblock copolymers polybutadiene-1,4-b-poly(ethylene oxide) (PB-b-PEO) using the same
water-soluble catalyst RhCl(TPPTS)3 . Different from their previous report using BR as
the substrate,71 the substrate PB-b-PEO itself is amphiphilic copolymer and can assemble
into micellar particles. However, in this PB-b-PEO emulsion system, the catalytic activity
is pretty low (TOF < 10 h−1) and the degree of hydrogenation that can be reached was
only 30 mol%. In contrast, the addition of cationic surfactant DTAC can increase the
hydrogenation rate drastically and a 100 mol% conversion was reached with a TOF =
32 h−1. This phenomenon can be explained that the added cationic DTAC has a pronounced
attraction force with the negatively charged RhCl(TPPTS)3 complex ions, which thus
will increase the concentration of Rh in the particles. Meanwhile, the existence of DTAC
created a smaller mixed emulsification system along with PB-b-PEO (hydrodynamic radius
= 41 nm), compared with the single micelle system formed by PB-b-PEO (hydrodynamic
radius = 58 nm). Finally, the authors pointed out the homogeneous essence of the latex
hydrogenation reaction using RhCl(TPPTS)3 .
However, relevant studies on bulk hydrogenation appear rare since Gilliom and co-
worker’s work, probably due to the difficulties resulting from the mass transfer and heat
transfer involved in the reactions.
Rempel et al. attempted to facilitate the transport of the catalyst inside the polymer
matrix through decreasing the dimensions of the substrate.73 In their experiments, NBR
solid was cut into small pieces with scissors (the average diameter of the NBR particles
being about 1 cm), and then the polymer pieces were physically mixed with the solid
RhCl(PPh3 )3 and TPP. The mixture was then transferred into a 5 mL flask that suspended
in a Parr high pressure reactor so that any possible overheating of sample could be avoided.
After the reactor was purged with hydrogen gas, the temperature and pressure were then
adjusted to 145◦ C and 3.45 MPa, respectively. The results showed that the NBR bulk could
be hydrogenated up to 95% conversion in 5 h.
Rempel and coworkers reported the bulk hydrogenation of NBR by taking advantage
of a supercritical fluid (SCF).74 The influence of scCO2 on the hydrogenation of NBR
entrapped with Wilkinson’s catalyst was studied in detail under various reaction conditions
characterized by reaction time, temperature, hydrogen pressure, loading level of catalyst,
and thickness of the polymer film. An 88.5% hydrogenation degree for a bulk NBR film
with a thickness of 0.0637 cm2/mg could be obtained at a RhCl(PPh3 )3 concentration of
0.046 wt% under a pressure of H2 and CO2 of 5.52 and 6.89 MPa, respectively, at 140◦ C
Hydrogenation of Nitrile Butadiene Rubber 215
for 3 h. It was observed that scCO2 has a strong swelling capacity for NBR and moves the
catalyst into or out of the NBR substrate effectively. Based on this phenomenon, the dis-
solution extent of RhCl(PPh3 )3 in scCO2 was estimated. It is found that high temperatures
and scCO2 densities, as well as in the presence of a co-solvent like acetone will enhance
the apparent solubility of catalyst. However, much more work, for example, the determi-
nation of the apparent solubility data of catalyst in scCO2 , is required in the upcoming
research.
where K (723) = 0.255 and K (970) = 2.3 are constants specific to the HNBR polymer.
Then, the relative amount of C C remaining in HNBR is:
Table 5
Characteristic peaks for NBR and HNBR FTIR spectrum.3
Figure 8 shows a typical hydrogenation process of NBR determined via FT-IR spectra.51
The peak intensity due to the olefinic double bonds at both 970 cm−1 and 917 cm−1
decreases gradually with the proceeding of the hydrogenation reaction. There is a new peak
at 723 cm−1 assigned to the (CH2 )n , n > 4, which is attributed to the oscillating H attached
to the saturated backbone CH2 CH2 . It can be seen in Fig. 8 that both the 970 cm−1
and 917 cm−1 peaks almost disappear after the hydrogenation reaction is completed.
Figure 8. A set of representative FT-IR spectra of 26 nm NBR nanoparticles with different extent of
hydrogenation. Experimental conditions of preparation of 26 nm NBR nanoparticles: AN = 5 mL,
BD = 15 mL, APS = 0.2 g, distilled water = 80 mL, T = 50◦ C, addition time = 200 min, aging
time = 8 h, concentration of surfactant is 40 mM. Experimental conditions of latex hydrogenation:
RhCl(PPh3 )3 /NBR is 1 wt%, RhCl(PPh3 )3 /TPP is 10 wt%, 600 rpm of agitation, 130◦ C, 6.89 MPa
of H2 . Reproduced from Wang et al.51 with permission from Wiley (Color figure available online).
Table 6
Chemical shifts of protons of different microstructures of NBR and HNBR.3
values were lowered with the reduction of the C C in the polymer backbone. The degree
of hydrogenation calculated from all three methods, IR, NMR, and iodometric analyses,
were comparable. Table 7 represents the degree of hydrogenation and corresponding iodine
values of the rubber samples.38
Table 7
Iodine value and degree of hydrogenation of nitrile rubber with different levels of bound
AN.38
Hydrogenation
Samples AN content, % I2 value degree, %
NBR 40 300 0
HNBR 40 156 48
HNBR 40 126 58
HNBR 40 90 70
HNBR 40 87 71
HNBR 40 72 76
HNBR 40 48 84
HNBR 40 42 86
HNBR 40 39 87
HNBR 40 15 95
HNBR 40 4 99
NBR 33 303 0
HNBR 33 25 92
NBR 51 295 0
HNBR 51 5 99
XNBR 34 308 0
HXNBR 34 35 89
BR 0 320 0
HBR 0 74 77
220 H. Wang et al.
Downloaded by [New York University] at 10:43 17 June 2015
Figure 10. 13C NMR spectra of acrylonitrile-butadiene copolymer (a) and hydrogenated acrylonitrile-
butadiene copolymer (b) (polymer solution in chloroform-d1 ). Reproduced from Mohammadi and
Rempel28 with permission from ACS.
commonly estimated as the temperature at the midpoint of the transition region from glassy
to rubbery, which were usually given by differential scanning calorimetry (DSC).
The superior flexibility retained by HNBR at low-temperatures is highly demanded
by the rubber parts that applied under rigorous working environment. The flexibility and
T g of HNBR are considered to be subject to the effects of both the level of bound AN
units and the length of the tetramethylene sequence in a polymer chain of HNBR.81–83 For
NBR without tetramethylene units, its T g was reported to increase linearly with an increase
of the molar fraction of AN units in a polymer chain, and thereby the flexibility and T g
of NBR can be very well controlled by the means of regulation of the levels of bound
AN of NBR. However, the same regulation practice is not effective for the production
of HNBR with the presence of tetramethylene units produced by the hydrogenation of
the BD of NBR, the crystallization of which leads to HNBR of poor flexibility in spite
of a low AN content.81 Some research work was carried out to explore the relationship
between the low-temperature flexibility of HNBR and the content of the chemical groups
Downloaded by [New York University] at 10:43 17 June 2015
excluding AN. Arsenault et al. investigated the relationship between the low-temperature
flexibility of HNBR and the molar ratio of ethylene units in it.82 Kobatake studied the
effects of the polymer structure of HNBR on its low-temperature flexibility, involving the
distribution of the monomer unit along a polymer chain and a higher order structure83
They reported that the length of the tetramethylene sequence contained in HNBR is the
key factor that governs the low-temperature flexibility of HNBR. Tetramethylene sequences
with more than five continuously linked tetramethylenes (TM-5) render crystallization upon
stretching, which results in the poor low-temperature flexibility of HNBR. Two effective
methods were proposed by Kobatake to reduce the TM-5 content so as to depress its
hydrogenated counterpart’s T g , as well as to improve its low-temperature flexibility. These
two modification methods were not directly used upon HNBR, but practiced in the stage
of NBR polymerization. The first approach is to copolymerize with a third monomer and
second one is to add mercaptans.83
The T g of HNBR is usually decreased with the increasing level of hydrogenation. T g
of the pure and hydrogenated nitrile rubbers are given in Table 8.38,83 T g of hydrogenated
rubber is lower than that of the pure polymer. The higher the degree of hydrogenation, the
Table 8
T g values of NBR and HNBR with different degree of hydrogenation and nitrile content.38,83
Hydrogenation
Elastomers AN content, % degree, % T g, ◦C
NBR 40 0 −18
HNBR 40 58 −19
HNBR 40 71 −20
HNBR 40 100 −24
NBR 33 0 −25
HNBR 33 92 −26
NBR 51 0 −11
HNBR 51 99 −18
Carboxylated nitrile (XNBR) 34 0 −23
HXNBR 34 89 −25
222 H. Wang et al.
Figure 11. WAXS pattern for HNBR stretched at 600% elongation: (a) HNBR-1 (49.9% ACN), (b)
HNBR-8 (42% ACN), and (c) HNBR-7 (16.35% ACN). Reproduced from Severe and White81 with
permission from Hüthig GmbH.
Downloaded by [New York University] at 10:43 17 June 2015
lower is the T g . This might be due to the chain flexibility upon the hydrogenation of C C
units. Hence, low temperature resistance improves on hydrogenation.
Severe and White investigated the phase transition behavior of HNBR and its tendency
to crystallize in the quiescent and stretched mode.81 It was found that the crystallinity can
be developed in HNBR through annealing and stretching NBR samples. The crystallinity at
high level of bound AN in HNBR was attributed to the alternating sequence of hydrogenated
BD and AN. With respect to the HNBR with low AN content, it was due to long sequences
of tetramethylene units. Severe and White confirmed the existence of crystallinity in HNBR
by using Wide-angle X-ray scattering (WAXS) pattern for three types of HNBR (Fig. 11).
Stretched HNBR with high level of bound AN such as 42% or 49.9 gives rise to the patterns
that contained diffused halo and three sharp reflections, which indicate that these samples
crystallized and oriented along the stretching direction (Fig. 11a and 11b). The diffraction
pattern for the stretched low AN content HNBR was diffused with no distinctive spot or
arc (Fig. 11c).
Scheme 7. Characteristic fragments of pyrolytic degradation for a segment of NBR chain with
adjoining AN and BD molecules after pyrolytic degradation. Reproduced from Hiltz87 with permission
from Elsevier.
degradation products comprising ethenylcyclohexene (the dimer of BD) and AN are shared
by all poly(butadiene-acrylonitrile) rubbers, but not exclusively by NBR.
Several studies have been reported for the purpose of the unambiguous identification of
NBR and HNBR. Alekseeva reported that the presence of AN, BD and ethenylcyclohexene
in the pyrolysis products can be considered as identification of NBR elastomer.91 Choi and
Han studied the pyrolysis behaviors of NBR with different levels of bound AN.86 Different
temperatures were practiced to pyrolyze NBRs with different AN contents: the temperatures
of 750, 670, 830, and 780◦ C were applied on the AN contents of 20.0, 33.4, 40.5, and
41.5 wt%, respectively. According to Choi and Han’s report, the characteristic pyrolysis
products of NBR involve BD, AN, benzene, cyclohexa-1,4-diene, 4-vinyl-1-cyclohexene,
cyclopent-3-ene-1-carbonitrile, cyanobenzene, and cyclohex-3-ene-1-carbonitrile.86 Hiltz
performed an extensive analysis focused on the pyrolytic degradation products related to
areas of NBR with adjoining AN and BD units (Scheme 7).87 By combining the pyrograms
of NBR, Hiltz summarized that the characteristic pyrolysate of NBR comprised of five
compounds with molecular weights of 93, 103 (cyanobenzene), 107, 119, and 160, which
can be employed to identify the NBR rubbers.
Kondo et al. investigated the microstructures of HNBRs with various degrees of hy-
drogenation by PyGC.93 They summarized the characteristic thermal degradation products
observed on the pyrolysates of NBR and hydrogenated NBR (Table 9). Typical mass spectral
data of the characteristic degradation products are shown in Table 10.93
224 H. Wang et al.
Table 9
Characteristic degradation products from hydrogenated NBR.93
Acrylonitrile A
Hydrocarbons
Mononitriles
Downloaded by [New York University] at 10:43 17 June 2015
EAE
BA
Dinitriles AEA
AEA
a
B = 1,4-butadiene unit; A = acrylonitrile; E = hydrogenated 1, 4-butadiene unit.
5. Recovery of Catalyst
5.1.2 Adsorption. Adsorption via solid materials probably is the most common practice
for recovery of organometallic catalysts from polymer solutions. A wide spectrum of
materials can be used as absorbents including silica, cross-linked polymers, inorganic
porous material, and hybrid materials of two or three of them. Generally, the process
involves passing the polymer solution through a column packed with the absorbent or
Hydrogenation of Nitrile Butadiene Rubber 225
Table 10
Typical mass spectral data of characteristic degradation products from hydrogenated NBR.93
mixing the absorbent with the solution followed by filtration or decanting. Panster et al.
used organo siloxane copolycondensates as absorbent for recovery of Rh and Ru catalysts
from HNBR solutions.99 A residue of Rh less than 5 ppm was reported by Panster et al.
with using the aforementioned absorbent. Additionally, ion-exchange resins modified with
different functional groups like thiourea were also reported with efficiency for recovery of
Rh catalysts from HNBR solution.100
5.1.3 Problems in Current Technology. Precipitation and adsorption as two principal cat-
alyst recovery techniques that employed in solution hydrogenation process suffer from
their respective drawbacks. The precipitation process carries high risk to contaminate the
final product if the additives that used for precipitation of the catalyst were not completely
removed, particularly for a process involving HNBR solution of high viscosity.99 The
226 H. Wang et al.
adsorption process involves a large amount of resin and related recycle/regeneration oper-
ation, which extends the time of a whole process and causes an increase of the expense.100
Furthermore, at the end of its recycle times, the resin is generally buried or dumped in an
open air, which becomes an environmental concern.
The technique of homogeneous catalyst recycle has made tremendous progress in the
past decade with the efforts from many chemists and engineers. Driven by the prominent
advantage of homogeneous catalyst in producing fine chemical products, a wide range of
creative solutions were proposed and investigated in their effectively and stably applications
for separation of homogeneous catalysts from products.101–107 These processes are still in
the stage of development and can be principally divided into two categories: (i) single-site
heterogenization of homogeneous catalyst, in which the catalyst was anchored on insoluble
or soluble macromolecules or supports to facilitate its separation by filtration in the end of
reaction; (ii) biphasic catalytic systems with exploitation of some versatile solvents such
as fluorous compounds, ionic liquids, and supercritical fluids, in which the transform of
Downloaded by [New York University] at 10:43 17 June 2015
the phase behavior of the reaction systems can be triggered by a series of factors, e.g.
temperature, pressure and small molecules etc. Although these processes showed certain
efficiency in recycling of homogeneous catalyst, none of them are fully developed for
industrial application as a stable process. The major challenges include the expensive
catalyst, ligand, solvent and the high catalyst leaching.
Figure 12. Reactions involved in the extraction process for separation of RhCl(PPh3 )3 from
HNBR by the means of CO2 -expanded methanol and N,N,N ,N ,N -pentamethyldiethylenetriamine
(PMDETA). Reproduced from Yang et al.116 with permission from Elsevier (Color figure available
online).
viscosity and increasing its permeability. The mass transfer resistance within HNBR was
thereby greatly reduced and most of the extraction was found to be completed within one
hour.116,117 After 9 h of operation at 80◦ C and 60 bar, only 59 ppm Rh remained in the
HNBR particles coagulated from the HNBR latex. Yang et al. indicated that the extraction
equilibrium was chemically determined.116 They reported that Wilkinson’s catalyst under-
goes dissociation of its TPP ligand at elevated temperatures and the CN group of HNBR
has a strong coordination interaction with these Wilkinson’s complexes RhCl(TPP)n (n ≤
2), which constitutes the central challenge involved in separation of catalyst from HNBR-
catalyst matrix. The extraction process and the challenges involved are elaborately illus-
trated in Fig. 12.116 Moreover, they pointed out that the existence of excessive TPP ligand
inside of HNBR plays a crucial role in releasing the Rh complexes from the CN group and
thus ensures their free diffusion within HNBR.
The application of CO2 expanded liquid technology in the recovery of organometallic
catalysts, that is, Wilkinson’s catalyst from polymers, that is, HNBR is still very promis-
ing, since this technology incorporates many advantages of the application of compressed
CO2 for polymer processing and homogeneous catalysis. As illustrated in Fig. 13, the
plasticization of the polymer by addition of scCO2 greatly improves the extraction pro-
cess by speeding up the overall mass transfer process at mild temperature and pressure.117
Meanwhile, CO2 expanded liquids can dissolve the objective organometallic catalysts at
mild temperature and pressure. CO2 expanded liquids are excellent solvents with tunable
solvation ability through manipulating the operational temperature and pressure, which fa-
cilitates the separation and purification for the recovery organometallic catalysts. Moreover,
with a large amount of replacement of organic solvents by environmentally benign CO2
provides this technology with green potential.
6. Applications of HNBR
Among the numerous hydrogenated elastomers, HNBR is a high performance specialty elas-
tomer which has excellent aging resistance and fluidic tolerance at high temperature.118–121
228 H. Wang et al.
Downloaded by [New York University] at 10:43 17 June 2015
Figure 13. Extraction process for recovering Wilkinson’s catalyst from HNBR-catalyst matrix via
CO2 -expanded methanol and N,N,N ,N ,N -pentamethyldiethylenetriamine (PMDETA). (a) extrac-
tion vessel; (b) one single polymer strip selected for the mass transfer investigation; (c) methanol
expanded by CO2 ; polymer plasticized by CO2 ; PMDETA absorbs on the surfaces of the polymer and
reacts with Wilkinson’s catalyst distributed on the polymer surfaces; (d) Wilkinson’s catalyst diffus-
ing from the interior of the polymer towards its surfaces; (e) Wilkinson’s catalyst coordinates with the
functional groups of the polymer and is retained in the polymer matrix; (f) separation of Wilkinson’s
catalyst terminated and the volume of polymer recovered after the CO2 release. Reproduced from
Yang et al.117 with permission from Elsevier (Color figure available online).
Figure 15. Sour environment test for tensile strength in liquid phase. Tensile strength in liquid phase
conditions: gas (H2 S 5%; CO2 20%; CH4 75%) and liquid (oil 95%; H2 O 4%; Amine B 1%) at 150◦ C
and 6.9 MPa. Reproduced from Hayashi121 with permission from Marcel Dekker.
230
Table 11
Common applications for HNBR.3, 118–126
Bulk
Figure 16. Major functional groups and their corresponding applications of HNBR.
applications. The seals made of HNBR are high demanded by oil field exploration and
processing under extreme high or low temperature and pressure conditions. The superior
mechanical properties at high temperature make HNBR also a dispensable material for
production of rolls for steel and paper mills.
7. Conclusions
Hydrogenation of unsaturated polymers to prepare the high performance specialty elas-
tomers and thermoplastics has been an important and interesting research area for both
industry partners and academic community. This article reviewed the fundamental knowl-
edge related to the hydrogenation of unsaturated rubbers and presented an overview of
the general features of catalysts that have been employed to hydrogenate NBR. Rh-based
complexes are the most important homogeneous catalysts for the highly efficient selec-
tive hydrogenation of polyolefins, and are particularly suitable for the hydrogenation of
NBR. Three reaction routes represented by the solution, latex, and bulk hydrogenations
were established to realize the selective hydrogenation of NBR. Comprehensive analysis
of the kinetics and mechanism of these three processes was provided. The conventional
techniques used for the characterization and analysis of NBR and HNBR microstructures
were well examined. Recent advances for catalyst recovery from polymer solutions and
latex systems were extensively discussed.
8. Future Outlook
(i) With respect to solution hydrogenation, some further work is suggested in or-
der to improve the productivity of the hydrogenation operation and to reduce the
production cost. In the cases of BR, SBR, SBS, and NBR, etc., hydrogenation
operation, it would be very important to develop a highly efficient reactor, which
can deal with highly viscous polymer fluids with superior transport capacity with
respect to increasing C C loading level in order to gain productivity. Enhancing
the transport behavior of a reactor for operation with high polymer concentration
232 H. Wang et al.
the latex hydrogenations process and the mechanism of transport of hydrogen and
catalyst across the interphase in the reaction system. Meanwhile, the exploration
of a highly efficient catalyst is demanded, in which a water soluble catalyst and
an in-situ activated catalyst provide promising research directions.
(iii) Another worthy pursuit involves a bulk hydrogenation operation assisted by super-
critical fluids, which would eliminate the impregnation operation of catalysts by
organic solvent and enhance mass transfer performance. In addition, considerable
insight into the kinetics and mechanism of bulk polymer hydrogenations should
be carried out, such as catalyst dissolution/diffusion in the polymer matrix.
(iv) It is very important to design and develop novel and efficient catalyst systems which
have high catalytic activities and production, as well as feasible and are economical
for recovery/recycle. Development of new heterogenization techniques for single-
site anchoring homogeneous catalyst is still promising for Wilkinson’s catalyst
recovery/recycle.
(v) Arising from the green hydrogenation processes based on latex form and bulk
form, the recovery of catalysts including both the oil soluble and water soluble
ones from the hydrogenated products with the aid of compressed CO2 holds great
potential to meet the requirements of industry.
(vi) It is highly desirable to extend new applications for hydrogenated polymers.
Acknowledgements
The authors gratefully acknowledge the support of Waterloo Institute for Nanotechnology
(WIN) and the NSERC Industrial Research Chair Program for the corresponding author
(G. L. Rempel).
Nomenclature
ACM Polyacrylate rubbers
AN Acrylonitrile
AEM Ethylene acrylate rubber
BD Butadiene
BR Polybutadiene rubber
CM Chlorinated polyethylene
CR Chloroprene rubber
Hydrogenation of Nitrile Butadiene Rubber 233
References
1. Konrad, E.; Tschunkur, E. “Rubber like masses from butadiene hydrocarbons and polymerizable
nitriles,” US Patent 1973000, Sep. 11, 1934; German Patent 658172, Apr. 26, 1930.
234 H. Wang et al.
2. Rempel, G.L.; Pan, Q.; Wu, J. “Homogeneous catalytic hydrogenation of polymers,” In Hand-
book of Homogeneous Hydrogenation; Johannes, G.de Vries.; Cornelis, J.E., Eds.; Wiley-VCH:
Weinheim, 2006; pp 547–579.
3. Singha, N.K.; Bhattacharjee, S.; Sivaram, S. “Hydrogenation of diene elastomers, their proper-
ties and applications: A critical review,” Rubber Chem. Technol. 1997, 70, 309–367.
4. McManus, N.T.; Rempel, G.L. “Chemical modification of polymers: catalytic hydrogenation
and related reactions,” J. Macromole. Sci. Part C-Polym. Rev. 1995, 35, 239–285.
5. Oppelt, D.; Schuster, H.; Thormer, J.; Braden, R. “Hydrogenation of polymers,” DE 2539132
A1, 1977.
6. Thoermer, J.; Mirza, J.; Szentivanyi, Z.; Obrecht, W. “Effect of crosslinking systems on pro-
cessability and performance profile of HNBR,” KGK-KAUT GUMMI KUNST 1988, 41,
1208–1214.
7. Wideman, L.G. “Process for hydrogenation of carbon-carbon double bonds in an unsaturated
polymer in latex form,” US Patent 4452950; 1984.
8. Lin, X.; Pan, Q.; Rempel, G.L. “Gel formation in diimide-hydrogenated polymers,” J. Appl.
Polym. Sci. 2005, 96, 1122–1125.
Downloaded by [New York University] at 10:43 17 June 2015
9. Belt, J.W.; Vermeulen, J.A. A.; Kostermann, M.; Dorrestijn, A. “Process for the hydrogenation
of a polymer composed of diene monomer units and nitrile group containing monomer units
and hydrogenated polymer,” WO 00/09568, 2000; granted by WIPO.
10. Lin, X.; Pan, Q.; Rempel, G.L. “Hydrogenation of nitrile butadiene rubber latex with diimide,”
Appl. Catal. A: Gene. 2004, 276, 123–128.
11. Zhou, S.; Bai, H.; Wang, J. “Hydrogenation of acrylonitrile-butadiene rubber latexes,” J. Appl.
Polym. Sci. 2004, 91, 2072–2078.
12. Belt, J.W.; Aagaard, O.M. “Process for the hydrogenation of a polymer composed of diene
monomer units and a nitrile group-containing monomer units,” WO 01/04162, 2001.
13. Xie, H.; Li, X.; Liu, X.; Guo, J. “Hydrogenation and neutralization of carboxylic styrene-
butadiene latex to form thermoplastic elastomer with excellent thermooxidation resistance,” J.
Appl. Polym. Sci. 2002, 83, 1375–1384.
14. Cowley, C.M.; King, J.G. “The hydrogen-cracking of rubber,” Rubber Chem. Technol. 1935, 8,
360–370.
15. Kobayasi, R.; Furihata, M.; Kajimoto, S. “Production of gasoline by means of hydrodepoly-
merization of rubber,” Rubber Chem. Technol. 1936, 9, 589–592.
16. van Leeuwen, P.W. N. M. Homogeneous Catalysis: Understanding The Art; Kluwer Academic
Publishers: Boston, 2004; pp 6–9.
17. Gilliom, L.R. “Catalytic hydrogenation of polymers in the bulk,” Macromolecules 1989, 22,
662–665.
18. Gilliom, L.R.; Honnell, K.G. “Observation of a reaction front in the bulk catalytic hydrogenation
of a polyolefin,” Macromolecules 1992, 25, 6066–6068.
19. Nagy, S.; Etherton, B.P.; Krishnamurti, R.; Tyrell, J.A. “Olefin polymerization catalysts con-
taining amine derivatives,” US Patent 6204216; 2001.
20. Pokasermsong, P.; Praserthdam, P. “Comparison of activity of ziegler-natta catalysts,” Eng. J.
2009, 13, 57–64.
21. Kaewarsa, P. “Polymerization of ethylene over the supported Ziegler-Natta and metallocene
catalysts on magnesium hydroxide and magnesiuum hydroxychloriede,” PhD Thesis, Khonkaen
University, Thailand, 2005.
22. Osborn, J.A.; Jardine, F.H.; Young, J.F.; Wilkinson, G. “The preparation and properties of
tris(triphenylphosphine)halogenorhodium(I) and some reactions thereof including catalytic ho-
mogeneous hydrogenation of olefins and acetylenes and their derivatives,” J. Chem. Soc. A
1966, 12, 1711–1732.
23. Ahmad, N.; Levison, J.J.; Robinson, S.D.; Uttley, M.F. “Complexes of ruthenium, osmium,
rhodium, and iridium containing hydride carbonyl, or nitrosyl ligands,” Inorg. Synth. 1974, 15,
45–64.
Hydrogenation of Nitrile Butadiene Rubber 235
24. Parent, J.S.; McManus, N.T.; Rempel, G.L. “RhCl(PPh3 )3 and RhH(PPh3 )4 catalyzed hydro-
genation of acrylonitrile-butadiene copolymers,” Ind. Eng. Chem. Res. 1996, 35, 4417–4423.
25. Charmondusit, K.; Prasassarakich, P.; McManus, N.T.; Rempel, G.L. “Hydrogenation of cis-1,4-
poly(isoprene) catalyzed by OsHCl(CO)(O2)(PCy3 )2 ,” J. Appl. Polym. Sci. 2003, 89, 142–152.
26. Bhattacharjee, S.; Bhowmick, A.K.; Avasthi, B.N. “Preparation of hydrogenated nitrile rubber
using palladium acetate catalyst: its characterization and kinetics,” J. Polym. Sci. Pol. Chem.
1992, 30, 471–484.
27. Bhattacharjee, S.; Rajagopalan, P.; Bhowmick, A.K.; Avasthi, B.N. “Selective hydrogenation
of olefinic bonds in styrene-isoprene-styrene triblock copolymer by palladium acetate catalyst,”
J. Appl. Polym. Sci. 1993, 49, 1971–1977.
28. Mohammadi, N.A.; Rempel, G.L. “Homogeneous selective catalytic hydrogenation of C C in
acrylonitrile-butadiene copolymer,” Macromolecules 1987, 20, 2362–2368.
29. Pan, Q.; Rempel, G.L. “Hydrogenation of Styrene-Butadiene Rubber Catalyzed by Ru(CH =
CHPh)Cl(CO)(PCy3 )2 ,” Macromol. Rapid Commun. 2004, 25, 843–847.
30. He, Y.; Daniels, E.S.; Klein, A.; El-Aasser, M.S. “Hydrogenation of styrene-butadiene rubber
(SBR) latexes,” J. Appl. Polym. Sci. 1997, 64, 2047–2056.
Downloaded by [New York University] at 10:43 17 June 2015
31. Lin, X.; Pan, Q.; Rempel, G.L. “Modeling and simulation of diimide hydrogenation of nitrile
butadiene rubber latex,” Ind. Eng. Chem. Res. 2006, 45, 1300–1306.
32. He, Y.; Daniels, E.S.; Klein, A.; El-Aasser, M.S. “Elastomeric films from structured latexes,”
J. Appl. Polym. Sci. 1997, 64, 1143–1152.
33. Garbisch, E.W.; Schildcrout, S.M.; Patterson, D.B.; Sprecher, C.M. “Strain effects. H. Diimide
reductions of olefins,” J. Am. Chem. Soc. 1965, 87, 2932–2944.
34. Pan, Q.; Rempel, G.L. “Numerical investigation of semibatch processes for hydrogenation of
diene-based polymers,” Ind. Eng. Chem. Res. 2000, 39, 277–284.
35. Pan, Q.; Rempel, G.L.; Ng, F.T. T. “Numerical investigation of continuous processes for catalytic
hydrogenation of nitrile butadiene rubber,” Polym. Eng. Sci. 2002, 42, 899–910.
36. Guo, X.; Rempel, G.L. “Catalytic hydrogenation of diene polymers. Part I. Kinetic analysis
and mechanistic studies on the hydrogenation of polybutadiene polymers in the presence of
chlorotris(triphenylphosphine)rhodium,” J. Mol. Catal. 1990, 63, 279–298.
37. Guo, X.; Parent, J.S.; Rempel, G.L. “Catalytic hydrogenation of diene polymers. Part II. Kinetic
analysis and mechanistic studies on the hydrogenation of styrene-butadiene copolymers in the
presence of RhCl(PPh3 )3 ,” J. Mol. Catal. 1992, 72, 193–208.
38. Bhattacharjee, S.; Bhowmick, A.K.; Avasthi, B.N. “High-pressure hydrogenation of nitrile
rubber: thermodynamics and kinetics,” Ind. Eng. Chem. Res. 1991, 30, 1086–1092.
39. Mao, T.F.; Rempel, G.L. “Catalytic hydrogenation of nitrile-butadiene copolymers by cationic
rhodium complexes,” J. Mol. Catal. A. Chem. 1998, 135, 121–132.
40. McManus, N.T.; Rempel, G.L. “Improvements in the hydrogenation of nitrile rubber using
Wilkinson’s catalyst,” Rubber Chem. Technol. 2008, 81, 227–243.
41. Hsu, K.; Wu, G.; Xu, R.; Yue, D.; Zhou, S. “Bimetallic complex catalyst systems, their prepa-
rations and application in the hydrogenation of unsaturated copolymers,” US Patent 6,084,033;
2000.
42. van Herk, M. Chemistry and Technology of Emulsion Polymerisation; Blackwell Publishing:
Oxford, 2005; pp 235–237.
43. Anastas, P.T.; Kirchhoff, M.M. “Origins, current status, and future challenges of green chem-
istry,” Acc. Chem. Res. 2002, 35, 686–694.
44. Papadogianakis, G.; Sheldon, R.A. “Catalytic conversions in water: environmentally attrac-
tive processes employing water soluble transition metal complexes,” New J. Chem. 1996, 20,
175–185.
45. Singha, N.K.; De, P.P.; Sivaram, S. “Homogeneous catalytic hydrogenation of natural rubber
using RhCl (PPh3 )3 ,” J. Appl. Polym. Sci. 1997, 66, 1647–1652.
46. Barrios, V.A. E.; Najera, R.H.; Petit, A.; Pla, F. “Selective hydrogenation of butadiene-styrene
copolymers using a Ziegler-Natta type catalyst 1. Kinetic study,” Eur. Polym. J. 2000, 36,
1817–1834.
236 H. Wang et al.
47. Rao, P.V. C.; Upadhyay, V.K.; Pillai, S.M. “Hydrogenation of polybutadienes catalyzed by
RuCl2 (PPh3 )3 and a structural study,” Eur. Polym. J. 2001, 37, 1159–1164.
48. Gehlsen, M.D.; Bates, F.S. “Heterogeneous catalytic hydrogenation of poly(styrene): thermo-
dynamics of poly(vinylcyc1ohexane) containing diblock copolymers,” Macromolecules 1993,
26, 4122–4127.
49. Hucul, D.A.; Hahn, S.F. “Catalytic hydrogenation of polystyrene,” Adv. Mater. 2000, 12,
1855–1858.
50. Widegren, J.A.; Finke, R.G. “A review of soluble transition-metal nanoclusters as arene hydro-
genation catalysts,” J. Mol. Catal. A: Chem. 2003, 191, 187–207.
51. Wang, H.; Pan, Q.; Rempel, G.L. J. “Diene-based polymer nanoparticles: preparation and direct
catalytic latex hydrogenation,” Polym. Sci. Pol. Chem. 2012, 50, 2098–2110.
52. Weinstein, A.H. “Elastomeric tetramethylene-ethylethylene-acrylonitrile copolymers,” Rubber
Chem. Technol. 1984, 57, 203–215.
53. Mahittikul, A.; Prasassarakich, P.; Rempel, G.L. “Hydrogenation of natural rubber latex in the
presence of OsHCl(CO)(O2 )(PCy3 )2 ,” J. Appl. Polym. Sci. 2006, 100, 640–655.
54. Mahittikul, A.; Prasassarakich, P.; Rempel, G.L. “Hydrogenation of natural rubber latex in the
Downloaded by [New York University] at 10:43 17 June 2015
70. MacLeod, S.; Rosso, R.J. “Hydrogenation of low molecular weight polymers in ionic liquids
and the effects of added salt,” Adv. Synth. Catal. 2003, 345, 568–571.
71. Kotzabasakis, V.; Hadjichristidis, N.; Papadogianakis, G. “Catalytic conversions in aqueous
media: Part 3. Biphasic hydrogenation of polybutadiene catalyzed by Rh/TPPTS complexes in
micellar systems,” J. Mol. Catal. A: Chem. 2009, 304, 95–100.
72. Kotzabasakis, V.; Georgopoulou, E.; Pitsikalis, M.; Hadjichristidis, N.; Papadogianakis, G.
“Catalytic conversions in aqueous media: a novel and efficient hydrogenation of polybutadiene-
1,4-block-poly(ethylene oxide) catalyzed by Rh/TPPTS complexes in mixed micellar nanore-
actors,” J. Mol. Catal. A: Chem. 2005, 231, 93–101.
73. Rempel, G.L.; Pan, Q.; Wu, J. “Organic solvent-free hydrogenation of diene-based polymers,”
EP 20060004541, 2006.
74. Li, G.; Pan, Q.; Rempel, G.L.; Ng, F.T. T. “Effect of supercritical CO2 on bulk hydrogenation
of nitrile butadiene rubber catalyzed by RhCl(PPh3 )3 ,” Macromol. Symp. 2003, 2004, 141–149.
75. Brück, D. “IR spectrometric determination of the proportions of acrylonitrile, butadiene and
hydrogenated butadiene in hydrogenated acrylonitrile-butadiene rubbers. Part 1. Principles,”
KGK-KAUT GUMMI KUNST 1989, 42, 107–110.
Downloaded by [New York University] at 10:43 17 June 2015
76. Brück, D. “IR spectrometric determination of the proportions of acrylonitrile, butadiene and hy-
drogenated butadiene in hydrogenated acrylonitrile-butadiene rubbers (HNBR). Part 2. Resid-
ual double bonds in commercial HNBR products,” KGK-KAUT GUMMI KUNST 1989, 42,
194–197.
77. Anachkov, M.P.; Stefanova, R.V.; Rakousky, S.K. “Proton NMR study of monomer sequence
distributions in some commercial acrylonitrile-butadiene copolymers,” Briti. Polym. J. 1989,
21, 429–432.
78. Suzuki, T.; Takegami, Y.; Furukava, J.; Kobayashi, E.; Aray, Y. “220-MHz NMR spectra of
butadiene and acrylonitrile alternating and random copolymers,” Polym. J. 1973, 4, 657–663.
79. Kuzay, P.; Kimmer, W. “Chain structural analysis of butadiene-acrylonitrile copolymers using
high-resolution nuclear magnetic resonance spectroscopy,” Plaste Kaut. 1971, 18, 743–744.
80. Santee, E.R.; Malotky, L.O.; Morton, M. “Analysis of chain unit structure of poly butadiene by
the HR-300 NMR spectrometer,” Rubber Chem. Technol. 1973, 46, 1156–1165.
81. Severe, G.; White, J.L. “Transition behavior of hydrogenated acrylonitrile-butadiene rubber,”
KGK-KAUT GUMMI KUNST 2002, 55, 144–148.
82. Arsenault, G.; Brown, T.; Jobe, I. “An approach to modern polymer development: Enhancement
of the service temperature range for hydrogenated nitrile-butadiene rubber (HNBR),” 145th
Meeting of the Rubber Division, ACS, Chicago, IL, 1994.
83. Kobatake, T.; Kodama, K.; Hayashi, S.; Yoshioka, A. “Improvement of low-temperature flexi-
bility of hydrogenated nitrile-butadiene rubber,” Rubber Chem. Technol. 1997, 70, 839–854.
84. Bhatacharjee, S.; Bhowmick, A.K.; Avasthi, B.N. “Hydrogenation of nitrile rubber using a new
homogeneous palladium(II) catalyst: Synthesis and characterization,” J. Appl. Polym. Sci. 1990,
41, 1357–1363.
85. The Hinrichs Lab in Organic Geochemistry at University of Bremen. http://www.
marum.de/en/HinrichsLab GC-MS.html (last accessed October 14, 2012).
86. Choi, S.S.; Han, D.H. “Pyrolysis behaviors of poly(acrylonitrile-co-butadiene) with differing
microstructures,” J. Anal. Appl. Pyrolys. 2007, 80, 53–60.
87. Hiltz, J.A. “Pyrolysis gas chromatography/mass spectrometry identification of poly(butadiene-
acrylonitrile) rubbers,” J. Anal. Appl. Pyrolys. 2000, 55, 135–150.
88. Wampler, T.P.; Bishea, G.A.; Simonsick, W.J. “Recent changes in automotive paint formula-
tion using pyrolysis-gas chromatography/mass spectrometry for identification,” J. Anal. Appl.
Pyrolys. 1997, 40–41, 79–89.
89. Hiltz, J.A. “Pyrolysis-gas chromatography/mass spectrometry identification of styrene cross-
linked polyester and vinyl ester resins,” J. Anal. Appl. Pyrolys. 1991, 22, 113–128.
90. Challinor, J.M. “A pyrolysis-derivatization-gas chromatography technique for the structural
elucidation of some synthetic polymers,” J. Anal. Appl. Pyrolys. 1989, 16, 323–333.
238 H. Wang et al.
91. Alekseeva, K.V. “Gas chromatographic identification of polymers using individual pyrolysis
products,” J. Anal. Appl. Pyrolys. 1980, 2, 19–34.
92. Hummel, D.O.; Dussel, H.J.; Czybulka, G.; Wenzel, N.; Holl, G. “Analytical pyrolysis of
copolymers,” Spectrochim. Acta 1985, 41A, 279–290.
93. Kondo, A.; Ohtani, H.; Kasugi, Y.; Tsuge, S.; Kubo, Y.; Asada, N.; Iwaki, H.; Yoshioka, A.
“Structural characterization of hydrogenated acrylonitrile-butadiene rubbers by pyrolysis gas
chromatography and infrared and nuclear magnetic resonance spectroscopy “, Macromolecules
1988, 21, 2918–2924.
94. Chamberlain, L.R.; Gibler, C.J. “Selective hydrogenation of conjugated diolefin polymers,” US
Patent 5,173,537; 1992.
95. Diaz, Z.; Gibler, C.J. “Removal of hydrogenation catalyst from polymer solutions by catalyzed
precipitation,” US Patent 5,212,285; 1993.
96. Gibler, C.J.; Austgen, D.M. R.; Parker, A. “Hydrogenation of polymers,” US Patent 6,177,521;
2001.
97. Wilkey, J.D. “Removal of hydrogenation catalyst from polymer solutions by treatment with
ammonia and carbon dioxide,” US Patent 6,207,795; 2001.
Downloaded by [New York University] at 10:43 17 June 2015
98. Hofman, H.; De Smet, H.J. A.; Villena, A.; Wirts, A.G. C. “Process for removing hydrogenation
catalyst residue from hydrogenated polymers,” US Patent 6,800,725; 2004.
99. Panster, P.; Wieland, S.; Buding, H.; Obrecht, W. “Recovery of hydrogenation catalysts from
solutions of hydrogenated nitrile rubber,” US Patent 5,403,566; 1995.
100. Nguyen, P.; Bender, H.; Arsenault, G.; Spadola, I.; Mersmann, F.J. “Process for removing
iron-and rhodium-containing catalyst residues from hydrogenated nitrile rubber,” US Patent
6,646,059; 2003.
101. Liu, Y. “Versatile routes for acrylonitrile butadiene rubber latex hydrogenation,” PhD Thesis,
University of Waterloo, Canada, 2012.
102. Schwab, E.; Mecking, S. “Immobilization of a catalytically active rhodium complex by elec-
trostatic interactions of multiply charged phosphine ligands with a soluble polyelectrolyte and
recovery by ultrafiltration,” Organometallics 2001, 20, 5504–5506.
103. Ropartz, L.; Haxton, K.J.; Foster, D.F.; Morris, R.E.; Slawin, A.M. Z.; Cole- Hamilton, D.J.
“Phosphine containing dendrimers for highly regioselective rhodium catalysed hydroformyla-
tion of alkenes: A positive ‘dendritic effect’” J. Chem. Soc. Dalton Trans. 2002, 23, 4323–4334.
104. Zheng, X.L.; Jiang, J.Y.; Liu, X.Z.; Jin, Z.L. “Thermoregulated phase transfer ligands and catal-
ysis. III. Aqueous/organic two-phase hydroformylation of higher olefins by thermoregulated
phase-transfer catalysis,” Catal. Today 1998, 44, 175–182.
105. Horváth, T.; Kiss, G.; Cook, R.A.; Bond, J.E.; Stevens, P.A.; Rábai, J.; Mozeleski, E.J. “Molec-
ular engineering in homogeneous catalysis: One-phase catalysis coupled with biphase catalyst
separation. The fluorous-soluble HRh(CO){P[CH2 CH2 (CF2 )5 CF3 ]3 }3 hydroformylation sys-
tem,” J. Am. Chem. Soc. 1998, 120, 3133–3143.
106. Koch, D.; Leitner, W. “Rhodium-catalyzed hydroformylation in supercritical carbon dioxide,”
J. Am. Chem. Soc. 1998, 120, 13398–13404.
107. Bronger, R.P. J.; Silva, S.M.; Kamer, P.C. J.; van Leeuwen, P.W. N. M. “A novel dicationic
phenoxaphosphino-modified Xantphos-type ligand-a unique ligand specifically designed for a
high activity, selectivity and recyclability,” Chem. Commun. 2002, 24, 3044–3045.
108. Curra, D.P. “Strategy-level separations in organic synthesis: From planning to practice,” Angew.
Chem. Ind. Ed. 1998, 37, 1174–1196.
109. Yoshida, J.; Itami, K. “Tag strategy for separation and recovery,” Chem. Rev. 2002, 102,
3693–3716.
110. Weiss, T.; Creutz, K. “Method for the hydrogenation of unsaturated polymers containing double
bonds,” US Patent 2008/0234437 A1; 2008.
111. Bhanage, B.M.; Ikushima, Y. “Multiphase catalysis using water-soluble metal complexes in
supercritical carbon dioxide,” Chem. Commun. 1999, 14, 1277–1278.
112. Bonilla, R.J.; James, B.R.; Jessop, P.G. “Colloid-catalysed arene hydrogenation in aque-
ous/supercritical fluid biphasic media,” Chem. Commun. 2000, 11, 941–942.
Hydrogenation of Nitrile Butadiene Rubber 239
113. Jacobson, G.B.; Lee, C.T. “Enhanced catalyst reactivity and separations using water/carbon
dioxide emulsions,” J. Am. Chem. Soc. 1999, 121, 11902–11903.
114. Matthew, J.C.; Kristi, L.H. “Water in supercritical carbon dioxide microemulsions: spectro-
scopic investigation of a new environment for aqueous inorganic chemistry,” J. Am. Chem. Soc.
1997, 119, 6399–6406
115. McCarthy, M.; Stemer, H.; Leitner, W. “Catalysis in inverted supercritical CO2 /aqueous biphasic
media,” Green Chem. 2002, 4, 501–504.
116. Yang, L.; Pan, Q.; Rempel, G.L. “Development of a green separation technique for recovery of
Wilkinson’s catalysts from bulk hydrogenated nitrile butadiene rubber,” Catal. Today 2012. (In
Press)
117. Yang, L.; Pan, Q.; Rempel, G.L. “Recovery of Wilkinson’s catalyst from polymer based matrix
using carbon dioxide expanded methanol,” J. Supercrit. Fluids 2012, 68, 104–112.
118. Mongiello, J. “Underhood elastomers market shifts to high performance materials,” Elastomer-
ics 1991, 123, 11–14.
119. Hoontrakul, P.; Szamosi1, J.; Tobing, S.D. “Application of nitrile butadiene rubber for flexible,
chemically protective coating,” Polym. Eng. Sci. 1988, 28, 1052–1055.
Downloaded by [New York University] at 10:43 17 June 2015
120. Threadingham, D.; Obrecht, W.; Wieder, W.; Wachholz, G.; Engehausen, R. “Rubber 3. Syn-
thetic rubbers, introduction and overview,” In Ullmann’s Encyclopedia of Industrial Chemistry,
6th Ed.; Wiley-VCH: Weinheim, 2011; pp 1–25.
121. Hayashi, S. “Nitrile and hydrogenated nitrile rubber,” In Handbook of Elastomers, 2nd Ed.;
Bhowmick, A.K.; Stephens, H.L., Eds.; Marcel Dekker: New York, 2001; pp 785–813.
122. Severe, G.; White, J.L. “Physical properties and blend miscibility of hydrogenated acrylonitrile-
butadiene rubber,” J. Appl. Polym. Sci. 2000, 78, 1521–1529.
123. Weissert, F.C. “Elastomer technology,” Ind. Eng. Chem. 1969, 61, 53–59.
124. Kubo, Y.; Mori, O.; Ohura, K.; Hisaki, H. “Adhesion between hydrogenated nitrile elastomers
and fibers is improved using hydrogenated acrylonitrile-butadiene latex,” Rubber Chem. Tech-
nol. 1990, 64, 8–18.
125. Donskoy, A.A. “Elastomeric heat-shielding materials for internal surfaces of missile engines,”
Int. J. Polym. Mater. 1996, 31, 215–236.
126. Ciesielski, A. An Introduction to Rubber Technology; Rapra Technology: UK, 1999; pp 18–19.