Sie sind auf Seite 1von 12

G Model

CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS


Catalysis Today xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Aqueous phase catalytic hydroformylation reactions of alkenes


Sumeet K. Sharma, Raksh V. Jasra ∗
Reliance Technology Group, Reliance Industries Limited, Vadodara 391 346, Gujarat, India

a r t i c l e i n f o a b s t r a c t

Article history: The present article provides an overview of the recent developments in the area of aqueous phase
Received 29 March 2014 hydroformylation of alkenes. The commercial application of aqueous phase hydroformylation reaction
Received in revised form 23 June 2014 is restricted to the lower carbon chain alkenes due to the solubility and mass transfer limitations of the
Accepted 30 July 2014
higher carbon chain alkenes in aqueous phase. The approaches which have been developed to improve
Available online xxx
the solubility of higher carbon chain length alkenes for aqueous phase hydroformylation reaction are
emphasized in this article.
Keywords:
© 2014 Elsevier B.V. All rights reserved.
Biphasic hydroformylation
Aqueous phase
Alkene
Aldehyde
Surfactant
Cyclodextrins

1. Introduction 2. Commercial significance

Hydroformylation or oxo reaction is a commercially important World production and consumption of hydroformylation (oxo)
reaction to produce aldehydes from the reaction of an alkene with chemicals is about 11 million metric tons per year with expected
syn-gas (mixture of hydrogen and carbon monoxide) in the pres- growth rate of 4.0% per year [6]. These oxo chemicals are used to
ence of a catalyst (Scheme 1). produce solvents, detergents, plasticizers, fragrances and interme-
The primary products of a hydroformylation reaction are linear diates for fine and specialty chemical industry.
and branched aldehydes with an additional carbon atom added to Propanal produced by ethylene hydroformylation and its
the reactant alkene. The obtained aldehydes can be converted into derivatives is used as a solvent, pesticide intermediates and print-
other commercially important chemicals via subsequent reactions, ing ink applications. Propanol is used as a precursor for glycol ether
like hydrogenation, condensation, amination, and oxidation. Gen- and surface coating applications. The other propanal derivatives,
erally, linear (n) aldehydes are more valuable than the branched sodium and calcium propionates finds application in the grain and
(iso) aldehydes, therefore, the n to iso ratio of aldehydes and food preservatives, 3,4-dichloropropionanilide as a herbicide and
the rate of formation are important parameters to be considered cellulose acetate propionate, as a plastic sheeting and molding pre-
in an industrial hydroformylation process. The alkene hydro- cursor.
formylation is an important reaction for C C bond formation and The market demand for C4 aldehydes which are synthesized by
functionalization of C C bonds with diverse functional groups to hydroformylation of propylene is higher among all oxo chemicals
produce fine chemicals [1–3]. The chiral compounds for the produc- due to its consumption for the production of 2-ethylhexanol and
tion of pharmacologically active molecules and agrochemicals can butanol. 2-Ethyhexanol is used for the production of dioctyl phtha-
be synthesized by asymmetric hydroformylation with controlled late, other plasticizers, coatings, adhesives, stabilizers, perfumery
enantio-, chemo- and regio-selectivity [4,5]. and specialty chemicals. 2-Ethylhexanol derivatives are used as
additives for diesel fuel to reduce emissions and for lube and mining
oil to improve its performance. n-Butanol is used for the production
of coating systems, cleaning fluids, herbicides, dyes, printing inks,
personal care products, pharmaceuticals, plasticizers, textiles and
lube additives.
∗ Corresponding author. Tel.: +91 265 6696313; fax: +91 265 6693934. Among C5 aldehydes, n-pentanal (valeraldehyde) is the fastest
E-mail addresses: sumeet.sharma@ril.com (S.K. Sharma), rakshvir.jasra@ril.com, growing oxo chemical, used for production of 2-propylheptanol.
rvjasra@gmail.com (R.V. Jasra). Valeraldehyde derivatives are used predominantly to make lube oil

http://dx.doi.org/10.1016/j.cattod.2014.07.059
0920-5861/© 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
2 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

Scheme 1. Hydroformylation reaction.

additives, which find application especially in automotive sector.


n-Valeric acid, which is prepared by hydroformylation of butene
followed by subsequent oxidation, is the basis of new ester type
lubricants for CFC-substitution in refrigeration systems. Volatile
esters of valeric acid are used in perfumes and cosmetics due to
their pleasant odor. The ethyl valerate and pentyl valerate are used
in food industries due to their fruity flavors.
C6 –C13 aldehydes are consumed for the production of fine and
perfumery chemicals, C6 –C13 alcohols for the synthesis of plasti-
cizers. C7–9 oxo acids which are the main derivatives of C7–9 oxo
aldehydes are used mainly to produce neopolyol esters. The 1-
heptanal which is produced by hydroformylation of 1-hexene is Fig. 2. Aqueous phase hydroformylation of alkene using HRh(CO)(TPPTS)3 complex.
a perfumery chemical and is also used for the production of lubri-
cants.
development [9–11]. The TPPTS ligand is highly soluble in water
Linear C12 –C18 aldehydes find their applications in the detergent
(c.a. 1.1 kg/L), mostly insoluble in common non-polar solvents and
industries to synthesize detergent grade alcohols. The detergent
sufficiently stable under reaction conditions [12]. The rhodium
grade alcohols can be converted into alcohol sulfates, ethoxylates,
complex of TPPTS ligand is a water soluble complex which can
alcohol ether sulfates and fatty amines for the diverse applications.
be separated easily from the reaction mixture after the comple-
tion of hydroformylation reaction in the aqueous phase. Despite all
3. Why aqueous phase hydroformylation the limitations discussed above, all the three generations of hydro-
formylation processes are still in commercial operation. Besides,
Typically, homogeneous catalysts comprising a transition metal cobalt and rhodium metals, ruthenium, iridium, platinum, palla-
and a ligand have been used for the hydroformylation of alkenes. dium and iron [13] have also been studied as alternative metals for
These transition metal complexes interact with carbon monoxide hydroformylation, but poor activity of these metals compared to
(CO) and hydrogen (H2 ) to form catalytically active metal car- cobalt and rhodium has restricted their application for the com-
bonyl hydride species. The first generation of the hydroformylation mercial hydroformylation process.
catalyst was exclusively based on the cobalt metal (Fig. 1). The sep- In the aqueous phase hydroformylation reaction, the active cat-
aration of products from reaction mixture, lower catalyst activity, alyst for the reaction is dissolved in water and remains in the
higher reaction temperature and pressure are the major limitations aqueous phase, whereas, reactants and products reside in the
of the first generation processes. organic phase. The mixture of aqueous and organic phases forms
The second generation of the hydroformylation processes com- two layers in the vessel and the reaction takes place at the inter-
bined developments for ligand modification and replacement of face as shown in Fig. 2. The separation of catalyst (water soluble)
the cobalt metal by rhodium. The rhodium metal modified by from the products mixture containing relatively non-polar organic
triphenylphosphine (PPh3 ) ligand is termed as Low Pressure Oxo compounds is carried out by simple decantation (phase separation).
(LPO) process [7]. The selectivity and n/iso ratio of aldehydes The aqueous phase containing the catalyst is recycled to the reactor
improved significantly with highly active and selective homo- and the organic phase is subjected for the purification of products.
geneous rhodium based complexes in LPO process. However, As compared to conventional homogeneous hydroformylation pro-
practical problems of separation of products from the reaction mix- cesses, further distillation is not required in the aqueous phase
ture, catalyst recovery and loss of expensive rhodium metal due hydroformylation process to separate the catalyst from reaction
to the lower thermal stability of the rhodium based complexes mixture.
continued in the second generation catalysts too. In the hydro-
formylation process, leaching of rhodium metal has significant
impact on the process economy. For example, leaching of 1 ppm 4. Commercial process for aqueous phase
rhodium per kg of product from a 400 kTA plant may result into hydroformylation of propylene and butene
the financial loss of several million euros [8]. These disadvantages
inspired the development of a water soluble ligand, trisodium salt The water soluble rhodium complex based on the TPPTS ligand
of tris(m-sulfonatophenyl)phosphine (TPPTS; P(m-C6 H4 SO3 Na)3 ) (1, Fig. 3) was commercialized in 1984 for the hydroformyla-
in the third generation (1980s) hydroformylation catalyst tion of propylene to produce C4 aldehydes at Oberhausen site
of Ruhrchemie AG (Ruhrchemie/Rhône-Poulenc (RCH/RP) pro-
cess) [11,14–19]. This process works at milder reaction conditions
(50 bar pressure and 120 ◦ C temperature) to give 95% propylene
conversion, 92–97% selectivity to the linear aldehyde with 99%
selectivity of total C4 aldehydes. The critical point in this pro-
cess is the production of TPPTS as it is highly unstable in the
presence of air/oxygen. In this process, TPPTS ligand is used
around 50 fold in the excess to rhodium metal to achieve higher
selectivity of n-aldehyde and to minimize rhodium leaching (less
Fig. 1. Development of hydroformylation catalysts. than 1 ppb rhodium leaching throughout catalyst batch life). The

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx 3

Fig. 3. Water soluble ligands.

constant activity of catalyst can be maintained for several years by (diphenylphosphino)-9,9-dimethylxanthene; 3, Fig. 3),
adding fresh TPPTS solution continuously or occasionally as per the NORBOS-Na (trisodium salt of 3,4-dimethyl-2,5,6-tris(p-
process requirement. In 1995, RCH/RP process was extended to the sulfonatophenyl)-1-phosphanorborna-2,5-diene; 4, Fig. 3), BINAS-
aqueous phase hydroformylation of raffinate-2 (generally 50–60% Na (octasodium salt of 2,2 -bis[(m-sulfonatodiphenylphosphino)-
n-butene, balance is cis/trans 2-butenes and butanes after separa- methyl]-4,4 ,8,8 -tetrasulfonato-1,1 -binaphthalene; 5,
tion of 1,3-butadiene) obtained from naphtha cracker to produce C5 Fig. 3) and BISBIS-Na (hexasodium salt of 2,2 -bis[(m-
aldehydes as 1-butene has acceptable solubility in water [18,20]. 1- sulfonatodiphenylphosphino)-methyl]-disulfonato-1,1 -biphenyl;
Butene rich cut and syn-gas are fed to the reactor in which rhodium 6, Fig. 3) were reported for aqueous phase hydroformylation of
and TPPTS ligand were present in the aqueous phase. The selectiv- alkenes [21–26].
ity to n-valeraldehyde was obtained up to 95% which is comparable The sulfonated diphosphane ligand (BISBIS-Na) gave excep-
to the selectivity of n-butanal produced by hydroformylation of tionally high n/iso ratio of butanals by aqueous phase rhodium
propylene. However, the rate of reaction for aqueous phase hydro- catalyzed hydroformylation of propylene [21]. Around 10–12 times
formylation of raffinate-2 was observed significantly lower than higher reactivity of the rhodium complex with BINAS-Na ligand was
the aqueous phase hydroformylation of propylene probably due observed than the conventional TPPTS ligand. The BINAS-Na ligand
to the lower solubility of butene in the aqueous phase than that also showed higher (98/2) n/iso ratio of aldehydes at lower ligand
of propylene. In aqueous phase hydroformylation of raffinate-2, to rhodium ratio (7/1) against 94/6 n/iso aldehydes ratio for TPPTS
slightly higher concentration of catalyst (150–500 ppm rhodium), ligand with 80/1 ligand to rhodium ratio (Fig. 4) [22,27,28]. Another
reaction temperature (120–130 ◦ C) and pressure (40–60 bar) were ligand, NORBOS-Na showed lower n/iso aldehydes ratio due to the
used to achieve the satisfactory level of valeraldehyde space time steric effect of ligand [22].
yields. In this process, 2-butenes remain almost unreactive due to The observed higher reactivity of the BINAS-Na ligand may be
steric and electronic nature of the TPPTS ligand. attributed to the electronic and steric effect of the ligand on the
hydroformylation reaction. The BINAS-Na ligand showed very high
4.1. Other water soluble ligands for aqueous phase n/iso ratio of aldehydes at lower ligand to rhodium ratio than the
hydroformylation of propylene conventional TPPTS ligand. The higher production cost and higher
decomposition rate of the BINAS-Na ligand are the main limiting
Apart from the TPPTS ligand, other water soluble ligands factors for its commercial application in the aqueous phase hydro-
(2–11, Fig. 3) such as SulfoXantPhos (2,7-bis(SO3 Na)-4,5-bis formylation reaction.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
4 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

Fig. 5. Rh-complex, ligand and ionic liquid as a solvent [35]. Reproduced with per-
mission from [35]. Copyright (2013) Elsevier.

cut. Alonso et al. carried out aqueous phase hydroformylation


of Venezuelan naphtha cut (42% C4 –C7 alkenes, 31% paraffins,
19% naphthenes, 7% aromatics) using HRh(CO)(TPPTS)3 com-
plex in the water [32]. Almost 86% conversion of alkenes was
Fig. 4. Reactivity of ligand for aqueous phase hydroformylation of propylene. Repro-
duced with permission from [22]. Copyright (1995) Elsevier. achieved in 200 h reaction time at 80 ◦ C, 800 psi syn-gas pres-
sure. The mono- and di-substituted alkenes present in naphtha
cut reacted faster (50% conversion in 24 h reaction time) than
4.2. Drawbacks of existing aqueous phase hydroformylation the sterically hindered alkenes. The concentration of paraffins,
processes naphthenes and aromatics remain constant during the reaction.
Very recently, aqueous phase hydroformylation of light naph-
The existing commercial aqueous phase hydroformylation pro- tha which contain around 61.5% alkenes (linear, branched, cis,
cess is restricted to the hydroformylation of propylene and trans, internal) was studied using HRh(CO)(TPPTS)3 complex [33].
1-butene due to their fair solubility in water. The rate of aqueous Around 95.4% conversion of alkenes with 95% yield of aldehydes
phase hydroformylation of alkenes depends upon the alkene sol- was obtained in 6 h at 70 ◦ C reaction temperature and 100 bar
ubility in water which makes process practicable at commercial syn-gas pressure in water-toluene mixed solvent. The n/iso ratio
scale. The higher carbon chain alkenes (>C6 ) have lower solu- of aldehydes was found to be 0.11 which is very low for rhodium
bility in the water which results into the lower reaction rates based catalyst, even significantly lower than the cobalt based
and mass transfer limitations. These factors make them unfa- catalyst.
vorable at commercial scale. To overcome these issues, other The sulfonated bidentate ligands are very selective to linear
approaches such as, use of co-solvents (mostly methanol, ethanol), aldehyde not only for hydroformylation of lower carbon chain
amphiphilic phosphine ligands, surfactants (cationic, anionic, dou- alkenes but also for the hydroformylation of higher carbon
ble long chain cationic), additives (cyclodextrins, activated carbon) chain length alkenes. Hamerla et al. achieved higher selectiv-
are reported. These approaches improved the interfacial area as ity to linear aldehyde (98/2) using SulfoXantPhos ligand with
well as the solubility of alkenes in the water during aqueous phase Rh(CO)2 (acac) in the presence of surfactant for hydroformylation
hydroformylation of alkenes. For example, solubility of 1-octene of 1-dodecene (TOF > 300 h−1 ) [26]. Interestingly, apart from
increased significantly (104 times) using mixture of 50% ethanol hydroformylation, SulfoXantPhos ligand was also found to be
and 50% water as a solvent instead of pure water [29]. Formation very active for aqueous phase hydroaminomethylation reaction
of the side products (e.g. acetals), decreased n/iso ratio of aldehy- in the presence of an imidazolium-based ionic liquid. Upto 95%
des, leaching of catalyst/metal are the major disadvantages of the yield of amines (n/iso ratio of amines = 78) were obtained with
use of co-solvents in the aqueous phase hydroformylation reaction. 16,000 h−1 TOF in the presence of piperidine as an aminating
By using a buffer solution of sodium carbonate and bicarbonate agent [34]. Recently, up to 98% conversion of 1-octene with 94%
(pH = 10), the acetals formation could be minimized [29]. Recently selectivity to the aldehydes (n/iso = 2.7) was reported for aqueous
Obrecht et al. compiled the alternative approaches which have hydroformylation of 1-octene using ionic complex, bis-[1-butyl-
been developed for the aqueous phase hydroformylation of alkenes 2-diphenylphosphanyl-3-methylimidazolium]tetrachloride-
[30]. In the present review, recent developments for the aqueous rhodium(III)hexafluorophosphate, 12, (Fig. 5) with ligand, 13,
phase catalytic hydroformylation of the higher carbon chain length (Fig. 5) and room temperature ionic liquid, 14, (Fig. 5) as a solvent
alkenes to overcome the solubility and mass transfer limitations [35].
are emphasized. The effect of steric and electronic factors of phosphine ligand
was studied by Wang et al. for aqueous phase hydroformylation
5. Rhodium metal complex catalyzed aqueous phase of 1-dodecene [36]. Two ligands, 2-MOTPPTS (tri(2-methoxyl-3-
hydroformylation of higher carbon chain (>C5 ) alkenes (sodium sulfonato)phenyl)phosphine; 8, Fig. 3), and 4-MOTPPTS
(tri(4-methoxyl-3-(sodium sulfonato)phenyl)phosphine; 9, Fig. 3)
The catalytic activity of HRh(CO)(TPPTS)3 and were synthesized and used for aqueous phase hydroformylation
HRh(CO)(TPPMS)3 (TPPMS; 2, Fig. 3) was studied by Baricelli of 1-dodecene [36]. Both ligands showed very low activity for
et al. for the hydroformylation of 1-hexene, 2,3-dimethyl-1- rhodium catalyzed hydroformylation of 1-dodecene due to the
butene, styrene and cyclohexene using mixture of water and poor solubility of 1-dodecene in the aqueous phase. Activity of 2-
n-heptane (1:1 by volume) as a solvent [31]. Higher catalytic MOTPPTS and 4-MOTPPTS was observed to improve significantly by
activity of HRh(CO)(TPPTS)3 and HRh(CO)(TPPMS)3 complex addition of CTAB. However, it is still lower than the conventional
was reported for the hydroformylation of 1-hexene followed by TPPTS ligand, perhaps, due to the presence of strong electron donat-
2,3-dimethyl-1-butene, styrene and cyclohexene at 80 ◦ C and ing methoxyl groups on the aromatic ring of MOTPPTS ligands.
54 atm syn-gas pressure (CO/H2 = 1/1). Interestingly, this approach The higher basicity of 2- and 4-MOTPPTS ligands is another rea-
was extended for the hydroformylation of mixed alkenes having son for lower activity of the MOTPPTS ligands than the TPPTS. The
medium carbon chain length (C4 –C7 ) present in the naphtha strong coordination of the MOTPPTS ligands with rhodium metal

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx 5

Fig. 6. Amphiphilic phosphine based ligands.

makes more difficult to dissociate one ligand molecule from the hydroformylation of 1-octene at 70 ◦ C and 40 bar syn-gas pres-
rhodium complex to form active catalyst species for the hydro- sure.
formylation reaction. In another study, rhodium polyethylene
glycolate complex was found highly active catalyst for the aqueous 5.1. Amphiphilic phosphines based ligands
phase hydroformylation of broad range of alkenes, C11 cut inter-
nal alkenes, cyclooctene, diisobutylene, 1-dodecene, 1-hexene, The amphiphilic phosphines (15–24, Fig. 6) are surface active
1-octene, styrene, ␣-methylstyrene and 2,4,4-trimethylpent-1-ene phosphines having a hydrophobic tail end and several hydrophilic
[37]. ends. These surface active phosphines work not only as a lig-
For the aqueous phase hydroformylation of terminal alkenes and for the rhodium or cobalt metal catalyzed hydroformylation
(4-pentenoate and acetoxystyrene), Boulanger et al. synthe- of alkenes but also as a surfactant to enhance the solubility of
sized two ligands based on the diphosphadiazacyclooctane alkenes in aqueous phase. Use of amphiphilic phosphine ligands
(10–11, Fig. 3) [38]. Complete conversion of 4-pentenoate with can show stronger stabilizing effect on metal species and effec-
99% aldehydes (n/iso = 1.1) selectivity was reported for hydro- tively decrease the interfacial resistances and phase separation
formylation of 4-pentenoate, whereas, lower conversion of time.
acetoxystyrene (83%) with 98% selectivity to aldehydes (n/iso = 0.5) The water soluble amphiphilic phosphine ligands (15–18,
was observed for Rh(CO)2 (acac) catalyzed hydroformylation of Fig. 6) were synthesized by Solsona et al. for the aqueous
acetoxystyrene using ligand (10) at 80 ◦ C and 50 bar syn-gas pres- phase hydroformylation of 1-octene in the mixture of water
sure. The ligand (11) showed lower activity and selectivity to and methanol as a solvent [40]. Significantly lower selectiv-
aldehydes for the hydroformylation of 4-pentenoate and ace- ity to the aldehydes was observed using ligands (15–18) than
toxystyrene. the commercial TPPTS ligand at 80 ◦ C and 1-octene to rhodium
The water soluble dendrimers were also used as ligands molar ratios 500 and 1000. The water soluble amphiphilic lig-
for the rhodium [RhCl(COD)]2 ; (COD = 1,5-cyclooctadiene) cat- and, sodium salt of sulfonated n-C12 H25 OC6 H4 P(C6 H4 -p-CH3 O)2
alyzed hydroformylation of 1-octene [39]. Hager et al. reported (DMOPPS; 19, Fig. 6) was also studied in combination of
the synthesis and activity of dendrimer based ligands (tris- RhCl(CO)(TPPTS)2 complex for the aqueous phase hydroformy-
2-(5-sulfonato-salicylaldimine-ethyl)-amine; diaminobutane-(5- lation of 1-octene, 1-decene, 1-dodecene, 1-tetradecene and
sulfonato-salicylaldimine)) and water soluble mononuclear 5- 1-hexadecene [41].
sulfonato propylsalicylaldimine Rh(I) complexes [39]. The den- The RhCl(CO)(TPPTS)2 complex with DMOPPS showed high
drimer base ligands were found to be more active (less rhodium catalytic activity for the hydroformylation of long chain alkenes
leaching) than the mononuclear based rhodium complexes for the (Table 1) due to the surface activity and micelle forming

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
6 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

Table 1
Aqueous phase hydroformylation using RhCl(CO)(TPPTS)2 complex with DMOPPS. Reproduced with permission from [41]. Copyright (2008) Elsevier.

Entry Alkene % Conversion % Selectivity n/iso t (h) TOF (h−1 )

Aldehydes Alkane iso-Alkenes

1 1-Octene 92.6 77.6 5.8 16.6 2.4 2 673.7


2 1-Decene 93.4 74.5 7.1 18.4 2.2 2 652.3
3 1-Dodecene 77.2 86.7 4.6 8.7 2.5 2 627.5
4 1-Tetradecene 97.1 83.9 4.1 12.0 2.6 22 69.4
5 1-Hexadecene 97.5 84.6 3.8 11.6 2.3 22 70.3

Reaction conditions: [RhCl(CO)(TPPTS)2 ] = 0.96 mmol/L, water = 2.5 mL, alkene/Rh molar ratio = 1875, DMOPPTS/Rh molar ratio = 6, syn-gas pressure = 3.0 MPa, T = 100 ◦ C,
600 rpm.

property of amphiphilic phosphine, DMOPPS, which assisted to Wang et al. studied the effect of TPPTS, HRh(CO)(TPPTS)2
stabilize the aqueous/organic phase boundary and enhanced the catalyst and alkene concentration on the critical micelle concen-
rate of hydroformylation reaction [41]. However, leaching of the tration (CMC) of cetyltrimethylammonium bromide (CTAB) [46].
rhodium from aqueous phase to organic phase was observed in The decrease in CMC and increase in the apparent molar mass
the range of 0.021–0.12 ppm during the reusability experiments of micelle was reported on the addition of TPPTS ligand and
for hydroformylation of 1-decene. HRh(CO)(TPPTS)2 complex in the aqueous solution of CTAB. The
Goedheijt et al. prepared the amphiphilic diphosphine lig- effect of reaction parameters such as, surfactant (CTAB) concen-
ands derived from the xantphos with surface active pendant tration, degree of emulsification on the rate of reaction, n/iso ratio
groups, -4-C6 H4 O(CH2 )nC6 H4 (SO3 Na)- (n = 0, 3, 6; 20, Fig. 6) [42]. of aldehydes was studied for the hydroformylation of 1-dodecene
The increased solubility of alkene in the aqueous phase was using RhCl(CO)(TPPTS)2 complex at 100 ◦ C and 1.1 MPa pressure
reported using these surface active groups, which resulted into [47]. The addition of TPPDS (di sulfonated) ligand along with the
the higher activity (upto 14 times higher than the xantphos) CTAB enhanced the n/iso ratio of aldehydes for RhCl(CO)(TPPTS)2
and reusability of catalyst for the aqueous phase hydroformyla- catalyzed hydroformylation of 1-dodecene due to the small steric
tion of 1-octene. Selectivity to linear aldehyde was reported in volume of the hydrophilic group [48].
the range of 94.7–96.9%. Other amphiphilic phosphine ligands The double long chain cationic surfactants (Fig. 8) were found to
(tris[p-(10-psulfonatophenyl-decyl)phenyl]phosphine, and chelat- be more effective than the other surfactants to improve the reaction
ing phosphine tetrasulfonated 2,2 -bis[di[p-(10-phenyldecyl)- rate for aqueous phase hydroformylation of long chain alkenes. The
phenyl]phosphinomethyl]-1,10-biphenyl) (21–22, Fig. 6) were double long chain cationic surfactants were synthesized by reflux-
found to be more active than the TPPTS for the aqueous phase ing the acetone or alcoholic solution of dimethyl alkyl amine and
hydroformylation of 1-octene and tetradecene [43]. The bio-ligands n-alkyl bromide [49].
based on amino acids (l-cysteine (24, Fig. 7), l-tryptophan (25, The conversion, selectivity to aldehydes, n/iso ratio of aldehydes
Fig. 7), l-methionine (26, Fig. 7)) and oligopeptides (glutathione, and turn over frequency (TOF) were observed to increase signifi-
l-cystine and vancomycin) were also studied for the hydroformy- cantly using the double long chain cationic surfactants as compared
lation reaction in aqueous medium [44]. to CTAB (Table 2) for the hydroformylation of 1-dodecene using
RhCl(CO)(TPPTS)2 as a catalyst with excess TPPTS ligand [49]. The
5.2. Surfactants for the aqueous phase hydroformylation work was extended to study the promotion effect of the cationic
Gemini and trimeric surfactants on RhCl(CO)(TPPTS)2 catalyzed
Surfactants (surface active agent; micellar systems) are studied hydroformylation of 1-decene [50].
for the aqueous phase hydroformylation reaction to increase the
interfacial area and solubility of an alkene in the aqueous phase. In
5.3. Mass transfer additives for aqueous phase hydroformylation
the aqueous phase hydroformylation of long chain alkenes, gen-
erally cationic surfactant increased the rate of reaction due to
Among the various approaches reported to increase solubility of
presence of anionic HRh(CO)(TPPTS)3 catalyst [45]. However, phase
higher carbon chain length alkenes in aqueous phase, use of chemi-
separation time may increase significantly due to the formation of
cally modified ␤-cyclodextrins is an attractive approach to enhance
highly stable emulsions. The enhanced hydroformylation rate is not
the rate of aqueous phase hydroformylation reaction. Cyclodex-
only due to increased solubility of an alkene in the aqueous phase
trins (Fig. 9) which are non-toxic, and inexpensive have been used
but also due to electrostatic interaction between positively charged
as mass transfer agents in the aqueous phase hydroformylation of
head group of the surfactant with negatively charged ligand (sul-
higher alkenes. Cyclodextrins work as carriers to transfer alkenes
fonate groups). The addition of nonionic surfactants to reaction
from the organic phase to the aqueous phase. Pioneer work has
mixture did not affect the rate of 1-dodecene hydroformylation
been done by Monflier et al. for the application of chemically modi-
whereas anionic surfactants showed adverse effect on hydroformy-
fied cyclodextrins in the aqueous phase hydroformylation reaction.
lation reaction [45].

Fig. 7. Bio-ligands based on amino acids for hydroformylation in aqueous medium. Fig. 8. Double long chain cationic surfactants. Reproduced with permission from
Reproduced with permission from [44]. Copyright (2007) Elsevier. [49]. Copyright (2006) Elsevier.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx 7

Table 2
Double long chain cationic surfactants for aqueous phase hydroformylation of 1-dodecene. Reproduced with permission from [49]. Copyright (2006) Elsevier.

Entry Surfactant % Conversion % Selectivity n/iso Ratio TOF (h−1 )

Aldehydes Alkane iso-Alkenes

1 C22 H45 N(CH3 )2 C16 H33 Br 83.3 96.9 2.6 0.5 3.7 1951.7
2 C22 H45 N(CH3 )2 C12 H25 Br 86.2 91.7 5.4 2.9 3.3 2019.7
3 C22 H45 N(CH3 )2 C8 H17 Br 85.9 93.0 3.4 3.6 3.6 2012.6
4 C16 H33 N(CH3 )2 C16 H33 Br 90.7 91.4 3.4 5.2 3.1 2125.1
5 C16 H33 N(CH3 )2 C12 H25 Br 94.5 93.0 4.2 2.8 3.3 2214.1
6 CTAB 20.9 80.4 10.0 9.6 2.8 489.7

Reaction conditions: 1-dodecene = 2 mL (9.0 mmol), [RhCl(CO)(TPPTS)2 ] = 9.6 × 10−4 mol/L, [TPPTS]/[Rh] = 18, H2 O = 4.0 mL, surfactant concentration = 5.0 mmol/L, syngas
pressure = 2.0 MPa, T = 100 ◦ C, t = 1 h at 400 rpm.

The use of chemically modified (methylated) ␤-cyclodextrins the randomly methylated ␤-cyclodextrins for Rh(acac)(CO)2 -TPPTS
for hydroformylation of 1-decene showed aldehyde selectivity catalyzed hydroformylation of 1-decene in the aqueous phase [60].
up to 95% with 100% conversion of 1-decene [51]. The forma- Additives other than cyclodextrins have also been studied
tion of alkene/cyclodextrin inclusion complex and solubility of to enhance the rate of hydroformylation reaction. The rate of
chemically modified ␤-cyclodextrins in the aqueous and organic hydroformylation reaction accelerated significantly by adding 1-
phases are the key factors to enhance the alkene conversion and octyl-3-methylimidazolium bromide as a mass transfer additive
selectivity to aldehydes. The formation of alkene/cyclodextrins for aqueous phase hydroformylation of 1-hexene, 1-octene and 1-
inclusion complex and cyclodextrins/ligand (m-TPPTC; tris(m- decene [61]. The nature of cationic head group of additive and its
carboxyphenyl)phosphane trilithium salt) interaction was studied alkyl side chain length has significant effect on the catalyst per-
in detail by NMR spectroscopy for the aqueous phase hydro- formance for the aqueous phase hydroformylation reaction [62].
formylation of alkenes [52,53]. The quantum chemistry calculation Chaudhari et al. reported that the organic solution (co-solvent) of
suggested that the ␤-cyclodextrins interact with either un- PPh3 can act as a promoter ligand for aqueous phase hydroformyla-
substituted or substituted (para and/or meta-position) sulfophenyl tion of 1-octene catalyzed by HRh(CO)(TPPTS)3 complex [63]. The
group of phosphines present in the ligand (TPPTS) [54]. Recently rate of reaction was observed to increase by a factor of 10–50 using
Tran et al. demonstrated the use of cyclodextrins for Rh(acac)(CO)2 PPh3 as a promoter ligand due to enhanced interfacial area of two
catalyzed hydroformylation of methyl 4-pentenoate (upto 99% con- phases (aqueous and organic). The activated carbon also has posi-
version and 98% selectivity to aldehydes) using TPPTS as a ligand tive effect as a mass transfer additive in the aqueous phase reaction.
[55]. The mixture of randomly methylated cyclodextrins of varied [64–68]. However, the exact role of activated carbon in the aqueous
size also has significant effect on the conversion of 1-tetradecene phase hydroformylation is not well understood but it does increases
due to the formation of ternary inclusion complexes between ran- the interfacial area due to its mesoporous and hydrophobic nature.
domly methylated cyclodextrins and alkene [56]. Neutral or ionic
nature of the modified ␤-cyclodextrins plays an important role for
6. Cobalt metal complex catalyzed aqueous phase
the aqueous phase hydroformylation reaction. Higher selectivity
hydroformylation
to linear aldehyde (n/iso ratio of aldehydes = 8.6; rate of reaction
increased 4 times) was achieved using ionic ␤-cyclodextrins in
Cobalt based homogeneous catalysts are cheaper than the
the stoichiometric quantities with ligand 23 (Fig. 6) for rhodium
rhodium complex and are still commercially used to produce
catalyzed hydroformylation of 1-decene [57]. Increase in the rate
detergent and plasticizer grade alcohols in the significant amount
of reaction, chemoselectivity and n/iso ratio of aldehydes was
(>2.5 million tons per annum) by hydroformylation of higher car-
reported for the use of cationic ␣-cyclodextrins as mass-transfer
bon chain length alkenes [13]. However, cobalt based catalysts
promoters for the Rh-TPPTS catalyzed hydroformylation reac-
are less active compared to the rhodium based complexes and
tion [58]. The ␤-cyclodextrins are useful mass transfer additive
require higher pressure and temperature. Guo et al. synthesized
for hydroformylation of terminal alkenes with enhanced reac-
Co2 (CO)6 (TPPTS)2 complex for aqueous phase hydroformylation of
tion rate using water-soluble, Rh2 (␮-StBu)2 (CO)2 (TPPTS)2 complex
1-hexene and used as a catalyst in the presence of excess TPPTS.
[59]. The modified ␤-cyclodextrins having methyl groups on the
[69]. Leaching of cobalt from the aqueous phase to the organic
secondary face and poly(ethylene oxide) chains on the primary
phase was observed during the reaction. Recently, Mika et al.
face were found to be more efficient mass transfer additives then
reported the recovery of cobalt from organic phase using aque-
ous solution of TPPTS ligand [70]. In another study, Beller et al.
reported the aqueous phase hydroformylation of 2-pentene (60%
E-2-pentene + 40% Z-2-pentene) using Co2 (CO)6 (TPPTS)2 complex
with 10 times excess TPPTS ligand in water and anisole mixture as
a solvent [71]. Upto 71.2% selectivity to the aldehydes was obtained
with 1.94 n/iso ratio of aldehydes at 190 ◦ C and 100 bar pressure.
The catalyst was recycled four times without any significant loss in
the activity. Parmar et al. prepared CoCl2 (TPPTS)2 complex for the
hydroformylation of 1-hexene in water as a solvent [72]. At 60 ◦ C,
63% conversion of 1-hexene with 83% selectivity to aldehydes (n/iso
ratio = 2.2) was reported under 9 MPa syn-gas pressure.

6.1. Effect of surfactants and mass transfer additives

The activity of cobalt complex, CoCl2 (TPPTS)2 was observed to


Fig. 9. Structure of ␤-cyclodextrins. increase in the presence of surfactant and co-solvents (methanol

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
8 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

Table 3
Aqueous phase hydroformylation of 1-octene and 1-decene. Reproduced with permission from Ref. [73]. Copyright (2008) Elsevier.

Entry
Alkene CTAB (mmol/L) TPPTS/Co molar ratio % Conversion % Selectivity n/iso Phases separation time (min)

Aldehydes iso-Alkenes Alkane

1 1-Octene 0 1 28 64 26 10 2.8 –
2 1-Octene 5.5 1 96 92 7 1 2.3 60
3 1-Octene 5.5 10 95 95 4 1 2.3 15
4 1-Decene 0 1 31 57 38 5 2.7 –
5 1-Decene 5.5 1 95 95 3 2 2.3 60
6 1-Decene 5.5 10 97 96 3 1 2.4 15

Reaction conditions: [CoCl2 (TPPTS)2 ] = 1.0 mmol/L, substrate/catalyst mol ratio = 700, T = 100 ◦ C, pressure = 8 MPa, solvent (water) = 25 mL, t = 5 h.

and ethanol) for the aqueous phase hydroformylation of 1-octene


and 1-decene [73]. Lower conversion of the alkene was reported in
the absence of surfactant. The concentration of surfactant plays an
important role not only on the alkene conversion but also on the
selectivity to aldehyde and phases separation time (Table 3).
On adding the CTAB, conversion of alkene, selectivity to alde-
hyde and phase separation time increased significantly [73]. The
phase separation time decreased from 60 min to 15 min using
excess TPPTS to the reaction mixture containing CoCl2 (TPPTS)2
and CTAB in the water. Leaching of the cobalt from aqueous
phase to organic phase was observed in CoCl2 (TPPTS)2 cat-
alyzed hydroformylation reaction. The leaching of cobalt was Scheme 2. Synthesis of florhydral using aqueous phase hydroformylation.
minimized by adding excess TPPTS ligand into the reaction
mixture. The catalyst was reused up to four times without
7. Synthesis of fine chemicals by aqueous phase
significant lose in alkene conversion, selectivity to aldehydes
hydroformylation
and n/iso ratio of aldehydes. Formation of alcohols up to 21%
was observed for Co/TPPTS catalyzed hydroformylation of 7-
Hydroformylation of functionalized alkenes is very use-
tetradecene (84% conversion) in microemulsion at 80 bar syn-gas
ful tool for the synthesis of fine and perfumery chemicals.
pressure and 160 ◦ C [74]. As observed for rhodium complex cat-
The 3-(3-isopropenylphenyl)butyraldehyde which is a fragrance
alyzed hydroformylation of alkenes, addition of cyclodextrins as
(Florhydral) precursor, was synthesized by rhodium catalyzed
mass transfer additives has significant effect on n/iso ratio of alde-
hydroformylation of m-diisopropenylbenzene (Scheme 2) in the
hydes for CoCl2 (TPPTS)2 catalyzed hydroformylation of 1-octene
mixture of water and toluene as a solvent using TPPTS and human
[75]. The conversion of alkene and selectivity to aldehydes was
serum albumin (HSA) as ligands [77].
reported almost similar to that when CTAB was used as a sur-
The [Rh(COD)Cl]2 with TPPTS ligand showed 81.3% conversion of
factant (without any cyclodextrins) but n/iso ratio of aldehydes
m-diisopropenylbenzene with 52.2% selectivity to the desired alde-
decreased from 2.3 observed for CTAB to 1.3 in case of randomly
hyde in water-toluene mixed solvent at 60 ◦ C and 70 atm syn-gas
methylated cyclodextrin, RAME-␤-CD as a mass transfer addi-
pressure. Florhydral can also be synthesized directly by the aque-
tive.
ous phase hydroformylation of 1-isopropyl-3-isopropenylbenzene.
In another study, TPPTS ligand was replaced by other water sol-
Almost complete conversion of 1-isopropyl-3-isopropenylbenzene
uble bulky ligand, trisulfonated tris(biphenyl)phosphine (BiphTS;
(99.6%) with 98.6% yield of Florhydral was achieved at 100 ◦ C
7, Fig. 3). The sodium salt of BiphTS has almost similar solubility
using Rh(CO)2 (acac)/TPPTS catalyst system. The HSA ligand showed
to TPPTS, more stable in air, easy to synthesize than TPPTS (98%
higher selectivity to aldehydes than TPPTS for aqueous phase
H2 SO4 used for synthesis of BiphTS against oleum (60%) for TPPTS).
Rh(CO)2 (acac) catalyzed hydroformylation of alkenes [78]. Another
Due to higher stability in air, 15–20% increased yields of BiphTS
fine chemical, 1,4-diacetoxy-2-formylbutane, an intermediate for
can be obtained than the TPPTS ligand [76]. The cobalt complex
the production of Vitamin A was synthesized by the selective
of BiphTS ligand, CoCl2 (BiphTS)2 showed lower conversion of 1-
aqueous phase hydroformylation of 1,4-diacetoxy-2-butene using
octene (23%) and 60% selectivity to aldehydes (n/iso = 2.5) than the
Rh-TPPTS catalyst (Scheme 3) [79,80].
CoCl2 (TPPTS)2 complex (29% conversion, 65% selectivity to alde-
Higher selectivity to the desired aldehyde and lower activity
hydes with n/iso = 2.8) under identical reaction conditions due to
of Rh/TPPTS was reported for aqueous phase hydroformylation of
strong coordination of BiphTS with metal (Co) than the TPPTS lig-
1,4-diacetoxy-2-butene than that of homogeneous Rh/PPh3 cata-
and [68]. The conversion of 1-octene increased to 95% with 86%
lyst [79]. Activity of Rh/TPPTS catalyst was observed to improve
selectivity to aldehyde (n/iso = 2.1) using 5.5 mol/L CTAB concen-
by adding the surfactant but it is still lower than the activity
tration with CoCl2 (BiphTS)2 catalyst. The conversion of 1-octene
observed with homogeneous Rh/PPh3 catalyst. Almost complete
decreased to 90% with 81% selectivity to aldehydes (n/iso = 1.3)
conversion of 1,4-diacetoxy-2-butene (99.9%) with 100% selectivity
using chemically modified cyclodextrin (RAME-␤-CD; 5 mol/L) as a
to 1,4-diacetoxy-2-formylbutane was achieved in the mixed sol-
mass transfer additive instead of CTAB [68].
vent (toluene and water) using [RhCl(COD)]2 /TPPTS catalyst [80].

Scheme 3. Synthesis of 1,4-diacetoxy-2-formylbutane by aqueous phase hydroformylation of 1,4-diacetoxy-2-butene.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx 9

Scheme 5. Aqueous phase hydroformylation of dicyclopentadiene.

Scheme 4. Aqueous phase hydroformylation of acrylates.


The role of water for aqueous phase hydroformylation of methyl
acrylate, ethyl acrylate, butyl acrylate, 2-ethoxyethyl acrylate and
The leaching of rhodium (less than 0.03%) was observed from aque- 2-ethylhexyl acrylate (Scheme 4) was studied by Fremy et al. [82].
ous phase to organic phase [80]. The disulfonated xantphos ligand Based on their results, it was observed that the water can also act
showed upto 82% yield of linear aldehyde for rhodium-catalyzed as a reactant or as a coordinating solvent to alter the catalytic cycle.
aqueous phase hydroformylation of 2-vinyl-5-methyl-1,3-dioxane The activity of RhCl(CO)(TPPTS)2 catalyst was studied for the
at 120 ◦ C [81]. aqueous phase hydroformylation of dienes (dicyclopentadiene;

Scheme 6. [Rh(CO)(Pz)(TPPMS)]2 catalyzed hydroformylation of allylbenzenes.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
10 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

Scheme 7. Aqueous phase hydroformylation of ␤-isophorone.

Scheme 8. Aqueous phase hydroformylation of various substituted olefins.

Scheme 5) in the presence of CTAB (1 mmol), which showed 99.3% Selectivity to the aldehydes was observed to increase in the
conversion of dicyclopentadiene with 95.2% selectivity to aldehy- presence of surfactant, cetyltrimethylammonium chloride (CTAC),
des at 140 ◦ C and 1.5 MPa syn-gas pressure [83]. although, phases separation time increased significantly [85]. Con-
Other than rhodium and cobalt metals, aqueous phase hydro- version of eugenol (98%) with 93% selectivity to aldehydes was
formylation of alkenes was also studied using ruthenium based reported using [Rh(CO)(Pz)(TPPMS)]2 complex and 1.02 mmol con-
catalysts. Baricelli et al. synthesized the ruthenium complex centration of CTAC at 80 ◦ C and 2.07 MPa syn-gas pressure. The
(Fig. 10; HRu(CO)(CH3 CN)(TPPTS)3 ]BF4 ) and characterized by FT-IR, rhodium complex showed higher catalytic activity for the hydro-
31 P NMR, etc. [84]. formylation reaction than the ruthenium complex. Among the
The activity of ruthenium complex (Fig. 10) for aque- studied ligands, TPPMS ligand was found to be more aldehyde selec-
ous phase hydroformylation of alkenes was reported in tive than the TPPTS for hydroformylation of eugenol, estragole,
the order of 1-hexene > allylbenzene > 2, 3-dimethyl-1- safrole and trans-anethole under studied experimental conditions.
butene > styrene > cyclohexene. The ruthenium complex showed The activity of ruthenium complex was observed in the order
sulphur (thiophene) tolerance limit upto 500 ppm concentra- of eugenol > estragole ∼ safrole  trans-anethole [85]. The rhodium
tion without having significant effect on the catalytic activity and ruthenium complexes with both of the studied ligands (TPPMS
which makes it a prominent catalyst for the hydroformylation and TPPTS) showed poor reusability of the catalyst for hydro-
of sulfur contaminated feed (e.g. alkenes rich naphtha cut). The formylation of eugenol. For aqueous phase hydroformylation of
aqueous phase hydroformylation approach is extended to the limonene, the addition of small amount of PPh3 (PPh3 :TPPTS ratio
synthesis of perfumery chemicals by hydroformylation of eugenol, of 0.05) showed improved conversion of limonene (12% in 21 h)
estragole, safrole and trans-anethole (Scheme 6) under moder- as compared to 1% without PPh3 using dinuclear water soluble
ate reaction conditions using binuclear water soluble rhodium [Rh2 (␮-St Bu)2 (CO)2 (TPPTS)2 ] complex [86]. The authors reported
[Rh(CO)(Pz)(L)]2 complexes and ruthenium complexes with ionic pale yellow color of organic phase which may be due to the leaching
liquid [HRu(CO)(CH3 CN)(L)3 ][BF4 ] (L = TPPMS and TPPTS) [85]. of catalyst from the aqueous phase to the organic phase on addition
of PPh3 .
4-Formyl-3,5,5-trimethylcyclohexan-1-one is an important
intermediate for the synthesis of ␦-damascone, a floral woody
perfume. Synthesis of 4-formyl-3,5,5-trimethylcyclohexan-1-one
was carried out by the hydroformylation of ␤-isophorone
using Rh/TPPTS catalyst (Scheme 7) [87]. The authors reported
91.7% conversion of ␤-isophorone in 72 h with 3.1% yield of
4-formyl-3,5,5-trimethylcyclohexan-1-one at 100 ◦ C, 100 atm syn-
gas pressure using mixture of water and toluene as a solvent. The
Fig. 10. Ruthenium complex for aqueous phase hydroformylation of alkenes [84]. reported lower yield of 4-formyl-3,5,5-trimethylcyclohexan-1-one

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx 11

Scheme 9. Aqueous phase hydroformylation of 2-benzyloxystyrene.

is due to the faster isomerization (21.4% yield) and hydrogenation [12] E.G. Kuntz, ChemTech 17 (1987) 570–575.
(43.4% yield) of ␤-isophorone. [13] J. Pospech, I. Fleischer, R. Franke, S. Buchholz, M. Beller, Angew. Chem. Int. Ed.
52 (2013) 2852–2872.
Paganelli et al. studied the hydroformylation of substituted [14] W.A. Herrmann, C.W. Kohlpaintner, Angew. Chem. Int. Ed. Engl. 32 (1993)
olefins (Scheme 8) for the synthesis of fine chemicals using 1524–1544.
Rh/TPPTS catalyst. [15] B. Cornils, W.A. Herrmann, R.W. Eckl, J. Mol. Catal. A: Chem. 116 (1997) 27–33.
[16] B. Cornils, Org. Process Res. Dev. 2 (1998) 121–127.
The Rh/TPPTS catalyst showed lower reactivity (59–99% conver- [17] P.J. Dyson, D.J. Ellis, T. Welton, Platinum Met. Rev. 42 (1998) 135–140.
sion with more than 99% selectivity to the aldehyde in 24–70 h) for [18] C.W. Kohlpaintner, R.W. Fischer, B. Cornils, Appl. Catal., A: Gen. 221 (2001)
the hydroformylation of 1,1-diarylethenes [88]. For aqueous phase 219–225.
[19] B. Cornils, W.A. Herrmann (Eds.), Aqueous-Phase Organometallic Catalysis, sec-
hydroformylation of 1,1 bis(p-fluoro-phenyl)-prop-2-en-1-ol, 99%
ond ed., Weley-VCH, Weinheim, 2004.

conversion with 92% selectivity to the aldehyde  reported
were [20] H. Bahrmann, C.D. Frohning, P. Heymanns, H. Kalbfell, P. Lappe, D. Peters, E.
at 100 ◦ C and 50 atm syn-gas pressure pCO = pH2 . In case of Wiebus, J. Mol. Catal. A: Chem. 116 (1997) 35–37.
[21] W.A. Herrmann, C.W. Kohlpaintner, J. Mol. Catal. 73 (1992) 191–201.
phenylvinylether, 98–99% conversion with 99% aldehyde selectiv-
[22] W.A. Herrmann, C.W. Kohlpaintner, R.B. Manetsberger, H. Bahrmann, H.
ity was achieved in 24 h reaction time. Kottmann, J. Mol. Catal. A: Chem. 97 (1995) 65–72.
Later this study was extended to the aqueous phase hydro- [23] H. Klein, R. Jackstell, M. Beller, Chem. Commun. (17) (2005) 2283–2285.
formylation of 2-benzyloxystyrene (Scheme 9) to produce [24] M. Scheuder Goedheijt, P.C.J. Kamer, P.W.N.M. van Leeuwen, J. Mol. Catal. A:
Chem. 134 (1998) 243–249.
pharmacological active compounds using TPPTS (1, Fig. 3) and Sul- [25] H. Nowothnick, A. Rost, T. Hamerla, R. Schomäcker, C. Muller, D. Vogt, Catal.
foXantPhos (3, Fig. 3) ligands [89]. Sci. Technol. 3 (2013) 600–605.
In case of TPPTS ligand, 99% conversion of 2-benzyloxystyrene [26] T. Hamerla, A. Rost, Y. Kasaka, R. Schomäcker, ChemCatChem 5 (2013)
1854–1862.
with 30% selectivity to linear aldehyde were achieved in 24 h reac- [27] B. Cornils, E.G. Kuntz, in: B. Cornils, W.A. Herrmann (Eds.), Aqueous-Phase
tion time, whereas, SulfoXantPhos showed very slow reaction rate Organometallic Catalysis, Wiley-VCH, Weinheim, 1998, p. 271.
(55.4% conversion of 2-benzyloxystyrene in 240 h) with higher [28] D.J. Cole-Hamilton, Science 299 (2003) 1702–1706.
[29] P. Purwanto, H. Delmas, Catal. Today 24 (1995) 135–140.
selectivity to the linear aldehyde (85.6%) than the TPPTS ligand. [30] L. Obrecht, P.C.J. Kamer, W. Laan, Catal. Sci. Technol. 3 (2013) 541–551.
[31] P.J. Baricelli, E. Lujano, M. Modrono, A.C. Marrero, Y.M. Garcia, A.
Fuentes, R.A. Sanchez-Delgado, J. Organomet. Chem. 689 (2004)
8. Future directions 3782–3792.
[32] M.M. Alonso, V. Guanipa, L.G. Melean, M. Rosales, A. Gonzalez, P.J. Baricelli,
Appl. Catal., A: Gen. 358 (2009) 211–214.
Subsequent to commercialization of aqueous phase hydro- [33] N.C. Kokkinos, E. Kazou, A. Lazaridou, C.E. Papadopoulos, N. Psaroudakis, K.
formylation of propylene and butene, numerous work have been Mertis, N. Nikolaou, Fuel 104 (2013) 275–283.
done for the development of aqueous phase catalytic process for [34] B. Hamers, P. Bäuerlein, C. Müller, D. Vogt, Adv. Synth. Catal. 350 (2008)
332–342.
hydroformylation of higher carbon chain length alkenes. The poor [35] S.J. Chen, Y.Y. Wang, W.M. Yao, X.L. Zhao, G.V. Thanh, Y. Liu, J. Mol. Catal. A:
solubility of higher carbon chain alkenes in water, mass trans- Chem. 378 (2013) 293–298.
fer limitations and lower reaction rates are main disadvantages [36] X. Wang, H.Y. Fu, X. Li, H. Chen, Catal. Commun. 5 (2004) 739–741.
[37] T. Borrmann, H.W. Roesky, U. Ritter, J. Mol. Catal. A: Chem. 153 (2000) 31–48.
of existing aqueous phase hydroformylation processes. Till today, [38] J. Boulanger, H. Bricout, S. Tilloy, A. Fihri, C. Len, F. Hapiot, E. Monflier, Catal.
none of the catalyst/process is commercialized for aqueous phase Commun. 29 (2012) 77–81.
hydroformylation of higher carbon chain length alkenes. There is [39] E.B. Hager, B.C.E. Makhubela, G.S. Smith, Hide affiliations, Dalton Trans. 41
(2012) 13927–13935 (Corresponding authors).
still need to develop commercially feasible low cost water soluble,
[40] A. Solsona, J. Suades, R. Mathieu, J. Organomet. Chem. 669 (2003) 172–181.
stable and selective catalyst as well as approaches to increase the [41] H. Fu, M. Li, J. Chen, R. Zhang, W. Jiang, M. Yuan, H. Chen, X. Li, J. Mol. Catal. A:
solubility of alkene in aqueous phase, simple phase separation in Chem. 292 (2008) 21–27.
minimum holding time and negligible leaching of rhodium/cobalt [42] M.S. Goedheijt, B.E. Hanson, J.N.H. Reek, P.C.J. Kamer, P.W.N.M. van Leeuwen, J.
Am. Chem. Soc. 122 (2000) 1650–1657.
metal from aqueous phase to organic phase. [43] B.E. Hanson, H. Ding, C.W. Kohlpaintner, Catal. Today 42 (1998) 421–429.
[44] S. Paganelli, M. Marchetti, M. Bianchin, C. Bertucci, J. Mol. Catal. A: Chem. 269
(2007) 234–239.
References [45] H. Chen, Y. Li, J. Chen, P. Cheng, Y. He, X. Li, J. Mol. Catal. A: Chem. 149 (1999)
1–6.
[1] R. Franke, D. Selent, A. Börner, Chem. Rev. 112 (2012) 5675–5732. [46] L. Wang, H. Chen, Yu-e He, Y. Li, M. Li, X. Li, Appl. Catal., A: Gen. 242 (2003)
[2] G.T. Whiteker, C.J. Cobley, Top. Organomet. Chem. 42 (2012) 35–46. 85–88.
[3] A. Behr, A.J. Vorholt, K.A. Ostrowski, T. Seidensticker, Green Chem. (2014), [47] C. Yang, X. Bi, Zai-Sha Mao, J. Mol. Catal. A: Chem. 187 (2002) 35–46.
http://dx.doi.org/10.1039/C3GC41960F (Advance article). [48] H. Chen, Y. Li, J. Chen, P. Cheng, X. Li, Catal. Today 74 (2002) 131–135.
[4] F. Agbossou, J.F. Carpentier, A. Mortreux, Chem. Rev. 95 (1995) 2485–2506. [49] H. Fu, M. Li, H. Chen, X. Li, J. Mol. Catal. A: Chem. 259 (2006) 156–160.
[5] B. Breit, Acc. Chem. Res. 36 (2003) 264–275. [50] H. Fu, M. Li, H. Mao, Q. Lin, M. Yuan, X. Li, H. Chen, Catal. Commun. 9 (2008)
[6] Chemical Insight and Forcasting: IHS Chemical. Oxo Chemicals: November 2012 1539–1544.
http://www.ihs.com/products/chemical/planning/ceh/oxo.aspx?pu=1&rd= [51] E. Monflier, G. Fremy, Y. Castanet, A. Mortreux, Angew. Chem. Int. Ed. 34 (1995)
chemihs. (Last accessed on 22.03.2014). 2269–2271.
[7] B. Cornils, W.A. Herrmann (Eds.), Applied Homogeneous Catalysis with [52] T. Mathivet, C. Meliet, Y. Castanet, A. Mortreux, L. Caron, S. Tilloy, E. Monflier,
Organometallic Compounds, 1, WILEY-VCH, Weinheim, 1996. J. Mol. Catal. A: Chem. 176 (2001) 105–116.
[8] K.D. Wiese, D. Obst, Catalytic carbonylation reactions, in: M. Beller (Ed.), Top- [53] S. Tilloy, E. Genin, F. Hapiot, D. Landy, S. Fourmentin, J.-P. Genêt, V. Michelet, E.
ics in Organometallic Chemistry, 18, Springer, Heidelberg, Germany, 2006, pp. Monflier, Adv. Synth. Catal. 348 (2006) 1547–1552.
1–33. [54] A. Sayede, M. Ferreira, H. Bricout, S. Tilloy, E. Monflier, J. Phys. Org. Chem. 24
[9] L. Gartner, B. Cornils, P. Lappe, EP Patent 0107006, 1983 to Ruhrchemie AG. (2011) 1129–1135.
[10] E.G. Kuntz, FR Patent 2314910 (1975) to Rhône-Poulenc Recherche. [55] D.N. Tran, F.X. Legrand, S. Menuel, H. Bricout, S. Tilloy, E. Monflier, Chem. Com-
[11] B. Cornils, E.G. Kuntz, J. Organomet. Chem. 502 (1995) 177–186. mun. 48 (2012) 753–755.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059
G Model
CATTOD-9237; No. of Pages 12 ARTICLE IN PRESS
12 S.K. Sharma, R.V. Jasra / Catalysis Today xxx (2014) xxx–xxx

[56] M. Ferreira, F.X. Legrand, C. Machut, H. Bricout, S. Tilloy, E. Monflier, Dalton [73] A.A. Dabbawala, D.U. Parmar, H.C. Bajaj, R.V. Jasra, J. Mol. Catal. A: Chem. 282
Trans. 41 (2012) 8643–8647. (2008) 99–106.
[57] M. Ferreira, H. Bricout, N. Azaroual, D. Landy, S. Tilloy, F. Hapiot, E. Monflier, [74] M. Haumann, H. Koch, R. Schomäcker, Catal. Today 79–80 (2003) 43–49.
Adv. Synth. Catal. 354 (2012) 1337–1346. [75] A.A. Dabbawala, J.N. Parmar, R.V. Jasra, H.C. Bajaj, E. Monflier, Catal. Commun.
[58] B. Sueur, L. Leclercq, M. Sauthier, Y. Castanet, A. Mortreux, H. Bricout, S. Tilloy, 10 (2009) 1808–1812.
E. Monflier, Chem. Eur. J. E (11) (2005) 6228–6236. [76] M. Ferreira, H. Bricout, F. Hapiot, A. Sayede, S. Tilloy, E. Monflier, ChemSusChem
[59] M. Dessoudeix, M. Urrutigoïty, P. Kalck, Eur. J. Inorg. Chem. 2001 (2001) 1 (2008) 631–636.
1797–1800. [77] S. Paganelli, A. Ciappa, M. Marchetti, A. Scrivanti, U. Matteoli, J. Mol. Catal. A:
[60] N. Badi, P. Guegan, F.-X. Legrand, L. Leclercq, S. Tilloy, E. Monflier, J. Mol. Catal. Chem. 247 (2006) 138–144.
A: Chem. 318 (2010) 8–14. [78] C. Bertucci, C. Botteghi, D. Giunta, M. Marchetti, S. Paganelli, Adv. Synth. Catal.
[61] S.L. Desset, D.J. Cole-Hamilton, D.F. Foster, Chem. Commun. (19) (2007) 344 (2002) 556–562.
1933–1935. [79] H.H.Y. Unveren, R. Schomacker, Catal. Lett. 102 (2005) 83–89.
[62] S.L. Desset, S.W. Reader, D.J. Cole-Hamilton, Green Chem. 11 (2009) 630–637. [80] R. Chansarkar, A.A. Kelkar, R.V. Chaudhari, Ind. Eng. Chem. Res. 46 (2007)
[63] R.V. Chaudhari, B.M. Bhanage, R.M. Deshpande, H. Delmas, Nature 373 (1995) 8629–8637.
501–503. [81] J. Ternel, J.L. Dubois, J.L. Couturier, E. Monflier, J.F. Carpentier, ChemCatChem 5
[64] N. Kania, B. Leger, S. Fourmentin, E. Monflier, A. Ponchel, Chem. Eur. J. 16 (2010) (2013) 1562–1569.
6138–6141. [82] G. Fremy, E. Monflier, Jean-Francois Carpentier, Y. Castanet, A. Mortreux, J. Mol.
[65] N. Kania, N. Gokulakrishnan, B. Leger, S. Fourmentin, E. Monflier, A. Ponchel, J. Catal. A: Chem. 129 (1998) 35–40.
Catal. 278 (2011) 208–218. [83] P. Xiaodong, Z. Yafen, Z. Limei, Y. Maolin, L. Ruixiang, F. Haiyan, C. Hua, Chin. J.
[66] N. Gokulakrishnan, N. Kania, B. Leger, C. Lancelot, D. Grosso, E. Monflier, A. Catal. 32 (2011) 566–571.
Ponchel, Carbon 49 (2011) 1290–1298. [84] P.J. Baricelli, K. Segovia, E. Lujano, M. Modrono-Alonso, F. Lopez-Linares, R.A.
[67] J. Boulanger, A. Ponchel, H. Bricout, F. Hapiot, E. Monflier, Eur. J. Lipid Sci. Sanchez-Delgado, J. Mol. Catal. A: Chem. 252 (2006) 70–75.
Technol. 114 (2012) 1439–1446. [85] L.G. Melean, M. Rodriguez, M. Romero, M.L. Alvarado, M. Rosales, P.J. Baricelli,
[68] A.A. Dabbawala, H.C. Bajaj, H. Bricout, E. Monflier, Appl. Catal., A: Gen. 413–414 Appl. Catal., A: Gen. 394 (2011) 117–123.
(2012) 273–279. [86] P. Kalck, M. Dessoudeix, S. Schwarz, J. Mol. Catal. A: Chem. 143 (1999) 41–48.
[69] I. Guo, B.E. Hanson, I. Toth, M.E. Davis, J. Organomet. Chem. 403 (1991) 221–227. [87] S. Paganelli, F. Battois, M. Marchetti, R. Lazzaroni, R. Settambolo, S. Rocchiccioli,
[70] L.T. Mika, L. Orha, E. Driessche, R. Garton, K. Zih-Perenyi, I.T. Horvath, J. Mol. Catal. A: Chem. 246 (2006) 195–199.
Organometallics 32 (2013) 5326–5332. [88] S. Paganelli, M. Zanchet, M. Marchetti, G. Mangano, J. Mol. Catal. A: Chem. 157
[71] M. Beller, J.G.E. Krauter, J. Mol. Catal. A: Chem. 143 (1999) 31–39. (2000) 1–8.
[72] D.U. Parmar, H.C. Bajaj, R.V. Jasra, B.M. Moros, V.A. Likholobov, J. Mol. Catal. A: [89] C. Botteghi, S. Paganelli, F. Moratti, M. Marchetti, R. Lazzaroni, R. Settambolo,
Chem. 211 (2004) 83–87. O. Piccolo, J. Mol. Catal. A: Chem. 200 (2003) 147–156.

Please cite this article in press as: S.K. Sharma, R.V. Jasra, Aqueous phase catalytic hydroformylation reactions of alkenes, Catal. Today
(2014), http://dx.doi.org/10.1016/j.cattod.2014.07.059

Das könnte Ihnen auch gefallen