You are on page 1of 19

J Am Oil Chem Soc

DOI 10.1007/s11746-016-2902-x

REVIEW

Catalysts for Fatty Alcohol Production from Renewable


Resources
Deepak S. Thakur1 · Arunabha Kundu1 

Received: 7 April 2016 / Revised: 13 September 2016 / Accepted: 14 September 2016


© AOCS 2016

Abstract  Fatty alcohols may be produced through the pro- Introduction


cessing of fatty acids or their esters derived from palm or
coconut oils. Fatty alcohol technology can be classified into Fatty alcohols are predominantly linear and monohydric
two categories (a) a slurry-phase process in which the fatty aliphatic alcohols with chain lengths between C8 and C22
acid or its ester is converted into alcohol using a powdered containing saturated or with one or more unsaturated C=C
catalyst and (b) fixed bed technology in which the fatty double bonds. The OH group is usually found in the pri-
acid or its ester is processed over a formed catalyst, e.g. mary position and provides a site for attaching strong
tablets or extrudates. Historically copper-based catalysts hydrophilic species. The hydrocarbon portion of the fatty
are employed for achieving ester hydrogenolysis. In recent alcohol molecule is hydrophobic, thereby providing the
years, studies were also focused on precious metal catalysts combination of hydrophobicity and hydrophilicity in the
for this process. This paper will critically review existing same molecule, as is required for a good surfactant.
literature pertaining to the catalysts that operate at diverse Commercially, fatty alcohols can be produced as linear
conditions, handle different feedstocks, and are compatible or branched, primary or secondary and saturated or unsatu-
with a variety of unit operations used by the fatty alcohol rated alcohol, depending upon the manufacturing process
manufacturers. There is an effort to develop environmen- and the raw materials. When natural products (fatty acid
tally friendly non-chrome copper catalyst in this process. methyl or wax esters) and ethylene are used as raw mate-
Some recent progress on bimetallic Cu-Fe catalyst and rials, the products are linear, primary and even-numbered
understanding Cu-Fe interaction has been reported. Current alcohols. The oleochemical raw material can yield both sat-
development on homogenous catalysts in this process was urated and unsaturated fatty alcohols. When olefins such as
carefully reviewed. The catalyst deactivation mechanism propylene are used as feed in oxo-process, branched alco-
has been investigated and effect of different impurities, hols with odd-numbered carbon chains can be produced.
mainly phosphorous, sulfur, chloride, water, glycerine and Fatty alcohols made by using renewable sources such
free fatty acid was thoroughly reviewed. as palm, palm kernel or coconut oil are termed as natural
alcohols. The carbon number in fatty alcohol depends on
Keywords  Oleochemistry · Hydrogenation · Processing the source of natural oil. For example, palm oil is rich in
technology C16–C18 while coconut and palm kernel oils are rich in C12–
C14. Detergent alcohols generally have a hydrocarbon chain
length of C12 minimum with at least 35 % linear chains.
Plasticizer alcohols generally have a chain length of C6–
C13 with a structure that is either linear or highly branched.
Synthetic alcohols are either produced by Ziegler process
* Deepak S. Thakur via ethylene polymerization with tri-ethyl aluminum fol-
ds.thakur@gmail.com
lowed by oxidation/hydrolysis, or by hydroformylation of
1
BASF R&D, 23800 Mercantile Road, Beachwood, OH propylene or its oligomers followed by hydrogenation of
44122, USA aldehydes.

13
J Am Oil Chem Soc

Fig. 1  Schematic of oleochemical process chemistry

Although the fatty alcohol technology has been prac-


ticed for a longtime, there has been continuous improve-
ment in the process and catalysts. Special emphasis was
given to increase the lifetime of the catalyst. Therefore the
deactivation studies with various poisoning components
were studied. The current review will focus on the catalysts
used for the production of natural alcohols. The term fatty
alcohol will be used in the foregoing sections when refer-
ring to natural alcohols. The other objective of this review
is to discuss critically the poisoning effect and catalyst
deactivation.

Fig. 2  Main applications of fatty alcohols [2–7]


Fatty Alcohol Market and Applications

With increasingly stringent environmental regulations and of global demand in 2014 with an expected growth of
the depletion of nonrenewable resources, the way has been 4.9 % from 2015 to 2022, which could be attributed to
paved for oleochemicals to enter the market and replace the availability of key raw materials in this region. The
the conventional petroleum-based products being used global fatty alcohols market is highly fragmented owing
currently. The factors driving growth in the market are: to the presence of several small and medium scale com-
high demand coming from consumer markets, easy avail- panies operating particularly in Asia Pacific and Central
ability of raw materials, and a growing market for green and South America. Major companies operating in the
chemicals. The oleochemicals market is expected to expe- global fatty alcohols market include Ecogreen Oleochemi-
rience huge growth due to the enormous potential offered cals, Emery Oleochemicals, BASF, Kuala Lumpur Kepong
by these drivers [1–7]. Recent studies [2, 3] suggest that (KLK), Oleon, Sasol, PT Muslim Mas, Saudi Kayan Pet-
the global fatty alcohols market demand was 2.4 million rochemical Company and Kao Chemicals.
tons in 2014 with C10–C14 fatty alcohols being the larg- Figure 1 presents at schematic of oleochemical technol-
est product segment with a market share of over 55 % ogy being practiced worldwide.
and is expected to reach 3.3 million tons by 2022. On the Applications of fatty alcohols depend upon the chain
other hand, C16–C22 fatty alcohols are expected to witness length. About 80 % of the mid-cut C12–C14 alcohols are
the highest growth rate of 4.5 % globally because of an used for production of detergents while the heavy-cut
increase in demand for lubricant bases and personal care C16–C18 alcohols are employed for the manufacture of
products. Asia Pacific is the largest consumer as well as health/personal care products. Figure 2 illustrates the main
producer of fatty alcohols. The region accounted for 35 % applications.

13
J Am Oil Chem Soc

Fatty Alcohol Chemistry Dehydration

The oldest method [8–10] for fatty alcohol production was 2RCH2 OH ⇋ R–CH2 − O –CH2 –R (7)
Fatty Alcohol − H2 O Ether
based on the reaction of triglycerides with sodium in the
presence of a reducing alcohol, such as amyl alcohol. A Dehydrogenation
disadvantage of this process was high processing costs due
RCH2 OH ⇋ R–CHO (8)
to the high price of sodium. Fatty Alcohol −H2 Carbonyl Compound
New processes were developed using hydrogen, which
is a cheaper and cleaner reducing agent than sodium. Com- A few studies have been conducted to elucidate the pos-
mercially, the fatty acids of palm, palm kernel or coconut sible mechanistic pathways. Norman [12] postulated that
oils are converted to esters by interesterification followed fatty acid ester hydrogenolysis proceeds through a hemiac-
by hydrogenolysis of esters to corresponding alcohols using etal intermediate:
copper-based catalysts [10–31]. However, the most com-
monly used catalyst appear to be based copper chromite.
Fatty acid (wax ester) and fatty methyl esters derived
from palm kernel or coconut oils are used as feeds. Fatty
acids are esterified with fatty alcohol at elevated tempera- (9)
tures to wax ester while triglycerides are transesterified
with methanol using sodium methylate solution or ion
exchange resin bed to form the methyl esters as shown by Commercially, both fatty acid wax ester and fatty acid
Eqs. (1) and (2), respectively. Commercially, the hydrog- methyl ester feeds are hydrogenated to fatty alcohol using
enolysis is carried out by using a copper-based catalyst slurry and fixed bed reactors. Table 1 gives the operating
(Eqs. 3 and 5). conditions used for each type of reactor to achieve opti-
mum performance.
Esterification/Transesterification
Catalytic Processes for Fatty Alcohol Production
RCOOH + RCH2 OH ⇋ RCOOR′ , (1)
Fatty Acid Fatty Alcohol Wax Ester
Earlier several groups in academia and industry [10–36]
where R′ = RCH2– were trying to develop catalytic processes for fatty alcohol
manufacture either from fatty acids or their corresponding
C3 H5 (OOCR)3 + 3CH3 OH ⇋ 3RCOOCH3 + C3 H5 (OH)3 esters. The pioneering work of Adkins et al. [11, 15, 16,
Triglyceride Methanol Methyl Ester Glycerin
(2) 23] documented an efficient way of producing fatty alco-
hols using copper chromite catalysts.
Ester Hydrogenolysis Initially, the fatty alcohol plants were installed by using
a high pressure continuous slurry process and copper chro-
RCOOR′ +H2 ⇋ 2RCH2 OH, (3) mite powder as catalyst. Procter & Gamble and the Kao
Wax Ester Fatty Alcohol
Soap used their own slurry methyl ester hydrogenation
where R′ = RCH2– process technology and in-house developed catalyst, while
Air Liquide (then Lurgi) licensed slurry fatty acid (via wax
RCOOCH3 +2H2 ⇋ RCH2 OH + CH3 OH ester) hydrogenation process with vendor supplied catalyst
Methyl Ester Fatty Alcohol Methanol (4)
to produce fatty alcohols. Air Liquide (then Lurgi) [37–39]
Interesterification developed a process that made it possible to hydrogenate
fatty acid by carrying out esterification to wax ester and
RCOOCH3 + RCH2 OH ⇋ RCOOR′ , (5) hydrogenation steps in one reactor using a copper chromite
Methyl Ester Fatty Alcohol Wax Ester
catalyst. The C=C double bonds present in the fatty acids
where R′ = RCH2– are hydrogenated to produce saturated fatty alcohols.
The undesired byproducts are formed during hydrogen- The slurry process is a continuous operation with fatty
olysis represented by Eqs. (5) to (7). ester feed, powder catalyst and hydrogen coming in contact
with each other under reaction conditions inside the reac-
Excessive Hydrogenation tor (Fig.3a). For the fatty acid or wax ester process(Fig.3b),
the wax ester is produced in situ by injecting fatty acid into
RCH2 OH −→ R–CH3 fatty alcohol and then hydrogenating to alcohol. The reac-
Fatty Alcohol +H2 Hydrocarbon (6) tor effluent goes through a few unit operations including

13
J Am Oil Chem Soc

Table 1  Slurry phase and fixed bed process conditions


Process conditions Slurry reactor Fixed bed
Liquid phase Gas phase

Pressure (bar) 250–300 250–300 40


Temperature (°C) 280 180–220 200–240
Feed Methyl or wax ester Methyl or wax ester Methyl ester
Catalyst shape Powder Tablets or extrudates Tablets or extrudates
H2/feed (molar ratio) 25–50 40–100 250 and above
Configuration Continuous recycle, liquid phase Down flow; trickle-bed, gas–liquid-solid Down flow; gas–solid

Fig. 3  Schematics of a methyl ester and b wax ester slurry processes [1]

gas–liquid separator to remove hydrogen, depressurizer to pressure process operates at slightly lower temperatures
reduce pressure to ambient conditions, and a continuous and lower hydrogen to feed ratio. In the vapor phase pro-
filtration system to remove the catalyst from fatty alco- cess, high temperatures and high hydrogen to feed ratios
hol. The catalyst withdrawal and addition rates (5–10 % are employed to keep the feed in gas phase (above dew
of total catalyst inventory in the reactor) are optimized for point). As expected, the capital costs are higher for the
maintaining desired conversion. The fresh catalyst is added high pressure process due to cost of reactor construc-
to the recycle catalyst slurry and gets activated in situ as tion material. Davy’s low pressure process is carried out
shown by Eq. (12). at lower pressure which reduces the capital expenses.
While Air Liquide (then Lurgi) was building slurry fatty However, the large volume of recycled hydrogen needed
alcohol plants, BASF [40–44] (then Henkel/Cognis) had to keep the reactant in the vapor phase requires a larger
developed in-house technology for a fixed-bed high-pres- compressor and could add some additional cost due to the
sure process to hydrogenate fatty acid methyl ester (FAME) separator and compressor.
to fatty alcohol using their own copper catalyst. In the There were a few attempts made to develop processes
1990s, the fixed bed fatty alcohol processes [1, 2, 3, 45–49] for direct one step hydrogenation of triglycerides to fatty
were introduced (Fig. 4) by Air-Liquide and Davy process alcohols [27, 28, 43]. Haidegger et al. [27] and Karo-
technologies. Air Liquide (Lurgi) technology is a high pres- lyi et al. [28] obtained the highest yields of fatty alcohol
sure liquid phase process operating at 250 bar while Davy by hydrogenation of triglycerides at 310–330 °C, 300 bar
process is a vapor phase process and operates at compara- pressure and 1–2 % catalyst concentration in a slurry pro-
tively lower pressure (40 bar), as shown in Table 1. cess, while Demmering et al. [43] patented a fixed bed pro-
The fixed bed operation is much simpler compared to cess operating at 200–230 °C and 250 bar hydrogen pres-
the slurry process. The catalyst is charged into the fixed sure using a CuZn catalyst. This process, however, yields
bed reactor either in pre-activated condition or is acti- a considerable amount of propylene glycol via glycerine
vated in situ before starting the feed. The trickle bed high hydrogenation.

13
J Am Oil Chem Soc

conversion is 65 % with the methanol to FAME mole ratio


of 5–15 and it is dropped to 35 % after the methanol/FAME
mole ratio is increased to 20. The selectivity is increased
with the increase of methanol/FAME mole ratio from 5 to
15 and it is maximum at 15 and then drops upon further
increasing the methanol/FAME mole ratio. As water helps
to produce hydrogen from methanol reforming, the pres-
ence of a small amount of water can enhance the conver-
sion of methyl laurate and the yield of lauryl alcohol. When
the molar ratio of water to methyl laurate was lower than 2,
the selectivity of lauryl alcohol did not change. The conver-
sion, yield, and selectivity were decreased when the molar
ratio of water to methyl laurate was higher than 2.
Most of the processes described above are based on
fatty ester conversion to fatty alcohol, Attempts were also
Fig. 4  Schematics of methyl and wax ester fixed bed processes [1] made to synthesize fatty alcohol by hydrogenation of fatty
acids using rhenium-ruthenium or ruthenium-tin cata-
Catalytic ester hydrogenolysis is normally performed lysts [57–70]. Broadbent et al. [57] conducted pioneering
in a multi-phase system. In such a system, because of low work to demonstrate the ability of ReO2 for hydrogenation
hydrogen solubility in the liquid substrate and mass trans- of carboxylic acids to alcohol. Following this approach,
fer resistance, the hydrogen concentration at the catalyst is Trivedi et al. [58, 59] showed that rhenium in combina-
low and limits the reaction rate. To overcome this limita- tion with palladium or ruthenium acts synergistically to
tion, supercritical solvents such as various alkanes (Van de hydrogenate fatty acids to fatty alcohols. Yoshino et al. [61]
Scheur et al. [50]) propane (van den Hark et al. [51, 52]), also confirmed these findings and reported that decanoic
butane (Brands et al. [53, 54]), n-pentane (Zhilong [55]), acid hydrogenation to decanol can be achieved by using a
and methanol (Liang et al. [56]) were used to make the combination Re–Os oxides and pre-activation in hydrogen.
reaction happen at a lower operating pressure. Table 2 reports the data generated at 130 °C and hydrogen
Zhilong [55] studied the effect of the molar ratio of pressure of 100 bar with dioxane as a solvent.
hydrogen to FAME, reaction temperature, space velocity Using a batch reactor, Tahara et al. [63] found that Ru–
and operating pressure in a fixed bed reactor with 15 g Cu– Sn catalysts were effective for the liquid-phase hydrogena-
Cr catalyst at supercritical conditions for palm oil fatty acid tion of the C=O group in carboxylic acids. Hydrogenation
methyl ester hydrogenolysis. In order to be in supercritical activity was affected by the kind of tin compounds used for
conditions, the typical weight ratio of the solvent to FAME the preparation of the Ru-Sn catalyst. Bis-tributyl tin oxide
is 90, whereas the temperature, H2/FAME molar ratio, ((Bu3Sn)2O), K2SnO3 and Na2SnO3 were found to be the
space velocity and operating pressure are 230–260 °C, 4–7, appropriate materials for the hydrogenation of the C=O
2–4 h−1 and 70–90 bar, respectively. group over the Ru–Sn catalysts. The C=O hydrogenation
As the hydrogen flow required to hydrogenate the activity for the hydrogenation of carboxylic acids increased
methyl ester is very low under the supercritical condition, with the increasing amount of Sn added whereas, the C=C
Liang et al. [56] took advantage of using methanol not hydrogenation activity decreased. Toba et al. [66] devel-
only as a supercritical solvent but also as an in situ hydro- oped Ru–Sn/Al2O3 catalysts for hydrogenation carboxylic
gen donor. Cu/ZnO/Al2O3 served as both a reforming and acids including C6–C18 fatty acids to corresponding alco-
hydrogenation catalyst for this purpose. The conversion is hols. The data shown in Table 3 compares the performance
very much dependent on the methanol concentration. The of these catalysts for various substrates. Rozmslowicz

Table 2  Hydrogenation of Precursors of catalysts Conversion (%) Products (%) Selectivity of alcohol (%)
decanoic acid with Re oxide
[61] Hydrocarbon Alcohol Ester

Re2O7 31.0 1.4 22.9 13.2 73.9


Re2O7 + OsO4 94.4 11.7 79.8 5.8 84.5
OsO4 3.4 1.1 1.6 1.3 47.1
Re2O7 + 5 % Ru/C 62.3 3.3 54.1 10.0 86.8
Re2O7 + 5 % Pd/C 8.7 2.0 4.7 3.2 54.0

13
J Am Oil Chem Soc

Table 3  Hydrogenation of fatty Substrate Carbon number H2 pressure (MPa) Acid conversion (%) Yield
acids over 2 wt% Ru/4.7 wt%
Sn catalysts [66] Alcohol Others

Valeric 5 5.6 52.0 50.4 1.6


Caproic 6 5.6 57.6 44.8 12.8
Caprylic 8 5.6 58.1 56.8 1.3
Capric 10 5.6 64.2 62.4 1.8
Lauric 12 5.6 65.0 64.4 0.6
Myristic 14 5.6 65.0 61.1 3.9
Palmitic 16 5.6 78.4 78.0 0.4
Stearic 18 5.6 77.2 76.3 0.9
Lauric 12 8.0 80.0 79.8 1.0
Myristic 14 8.0 84.9 83.5 1.4
Palmitic 16 8.0 82.2 75.3 6.8
Stearic 18 8.0 82.7 79.8 2.9

et al. [70] studied selective hydrogenation of stearic acid The fresh copper chromite was found to consist of CuO
over 4 % ReOx/TiO2 at temperatures of 180–210 °C and and CuCr2O4. In addition to the reactions (1)–(5), the fol-
20–40 bar with selectivity of 93 %. The use of reduc- lowing reactions occur during ester hydrogenolysis using
ible oxide support, such as TiO2 instead of alumina has copper catalysts:
increased the activity and selectivity due to Ru–Ti metal
support interaction. Cu2+ O / Support + H2 → Cu0 /Support +H2 O (Desired Reaction)
Catalyst Active Site
Although there have been more studies on the use of (12)
rhenium-ruthenium catalysts [63–70] for carboxylic acid
hydrogenation, these catalyst systems have not yet been
used in the fatty alcohol industry. This may be due to the Cu2+ O /Support + RCOOH → (RCOO)2 Cu2+ +H2 O
Catalyst Free Fatty Acid Copper Soap
catalyst price due to use of a precious metal and corro- (13)
(Undesired Reaction)
sion-resistant materials of construction for the process
requirements.
(RCOO)2 Cu2+ + H2 → 2RCH2 OH + Cu0
Catalyst Development for Slurry Process
 
Undesired Cu0 Sintering (14)

Initially, copper chromite had been the catalyst of choice, Finely dispersed metallic copper, formed by the reduc-
prepared by several ways described in the literature [11, tion of CuO (Eq. 12), was found to be the active site for
13–16, 23, 31–36]. However, the method developed by ester hydrogenolysis.
Adkins et al. [11, 15, 16] gave the most satisfactory cata- Norman [12] used copper carbonate supported on a sil-
lyst. Their procedure involves precipitation of an interme- ica carrier for coconut oil hydrogenation at 310–315 °C and
diate, which is a double hydroxide with copper and chro- pressures ranging from 120 to 500 bar. No hemiacetal was
mium with ammoniacal specie—commonly referred to isolated but the ester hydrogenation was found to depend
as Adkins’ complex—followed by its decomposition to a upon hydrogen pressure. Schrauth et al. [11]. established a
mixture of copper oxide and copper chromite. The reaction high pressure process operating at 320 °C and 300 bar pres-
schemes are illustrated by following Eqs. (10) and (11): sure using copper catalysts.
Church and Abdel-Gelil [26] studied the effect of copper
2 Cu(NO3 )2 + (NH4 )2 Cr2 O7 + 4NH4 OH chromite catalyst concentration, and other process param-
→ 2 NH4 CuCrO4 (OH) +4NH4 NO3 + H2 O (10) eters such as temperature, pressure and agitation rate on
Adkin’s Complex
the rate of methyl laurate hydrogenation to lauryl alcohol.
During decomposition of Adkins’ complex, the ammo- Conversions of 90 % or higher were achieved at tempera-
nia splits into nitrogen and hydrogen. The latter reduces tures and pressures ranging from 280 to 300 °C and 200–
Cr6+ to Cr3+ species which is an exothermic reaction. 250 bar, respectively with 4 % catalyst concentration. Cata-
lyst was recycled several times to determine reusability of
NH4 CuCrO4 (OH) → CuO · CuCr2 O4 + N2 + 5H2 O (11) the catalyst for a continuous operation. Pantulu and Achaya

13
J Am Oil Chem Soc

Table 4  Effect of calcination temperature on physical properties of CuCr catalyst [79]


Catalyst number Calcination temperature (°C) BET area (m2/g) Pore volume (mL/g) Pore size (Å) Catalyst diameter (μ)

1 399 44.8 0.31 276 1.66


2 427 35.9 0.24 268 1.58
3 460 25.8 0.19 296 1.28
4 516 18.5 0.16 344 1.23
5 538 13.4 0.09 280 1.02

[29] demonstrated that fatty alcohol can be produced by ester hydrogenolysis, the presence of feed impurities, such
hydrogenating fatty acids and fatty acid methyl esters using as glycerine, water etc., can adversely affect the catalyst
copper soaps as the catalyst. performance.
In the same time period, Muttzall [31, 32] studied the
kinetics of high pressure hydrogenolysis of wax esters to Activity/Selectivity
fatty alcohols and found that ester hydrogenolysis is an
equilibrium reaction. For optimum activity/selectivity, the copper chromite
(unpromoted or promoted by Ba or Mn) has met the above
Keq
RCOOR′ + H2 ⇆ 2RCH2 OH (15) requirements and has been the catalyst of choice for con-
tinuous slurry operation for several decades [11, 16,
He reported that the equilibrium constant (Keq) is 71–80]. Rieke et al. [79, 80] found that both activity and
dependent upon temperature and slightly dependent on selectivity correlate well with the crystallinity of the cop-
the molecular weight of the fatty alcohol. The rate of ester per chromite surface; they increase with decreasing crys-
hydrogenation to fatty alcohol depends linearly on the reac- tallinity or increase in surface area. Table 4 provides calci-
tion rate constant, catalyst and ester concentration, and on nation temperatures and surface area for the catalysts used
the square of the hydrogen pressure. in this study and Fig. (5a, b)illustrates their activity and
Boerma [35] has shown that the pH range for maximiz- selectivity respectively.
ing precipitation of copper and chromium hydroxide with The main goal of their laboratory investigation was to
ammonium hydroxide is very narrow (around 5.3–5.8). develop a superior copper chromite catalyst for the slurry-
Beyond this range both species form soluble salts and end phase process. Two copper chromite catalysts, prepared by
up in the filtrate. According to Boerma [35], the presence different procedures, were tested for methyl ester hydrog-
of barium in the copper chromite catalyst helps prolong the enolysis activity, reusability, and filtration characteristics.
catalyst life. The first sample (CuCr-I) was prepared by simultaneous
precipitation at constant pH described in the patents granted
Challenges for Slurry Phase Catalyst Designer to Thakur et al. [77, 78], while the second sample (Cu–
Cr-II) was prepared by sequential precipitation in which pH
In the slurry process, the overall production rate depends was varied with time. The former sample had narrow par-
upon catalyst activity and its separation/filtration rates. ticle and pore size distribution while the latter had a broad
While the reaction rate depends upon the catalyst composi- particle and pore size distribution. The reaction was carried
tion and surface properties, the settling or separation rate out in a batch autoclave at 280 °C and 140–210 bar hydro-
for continuous filtration is dictated by the physical proper- gen pressure. The reaction rates were calculated by assum-
ties such as particle size and skeletal liquid filled density. ing a kinetic mechanism that was first-order in methyl ester
Catalyst suffers severe attrition during the course of the concentration. The catalyst prepared by procedures devel-
process, which also results in slower separation rate. oped by Thakur et al. [77, 78] (CuCr-I; Fig.6) was 30 %
While designing a catalyst for continuous slurry-phase more active (Fig.6a), filtered faster (Fig.6c) and maintained
systems, in addition to optimizing the activity, the catalyst activity (Fig.6b) for several uses than the catalyst (CuCr-II)
formulator is faced with the challenge of improving the with the broader particle size distribution. X-ray photoelec-
factors, namely: in situ activation, poison resistance, attri- tron spectroscopy data showed higher surface copper con-
tion resistance and ease of separation of the catalyst from centrations for the former catalyst which may be attributed
the product, which can significantly impact the overall effi- for higher activity.
ciency of the process. In fatty acid process, the acid resist- Prasad et al. [81, 82] have summarized several prepa-
ance of the catalyst plays an important role in determining ration techniques for synthesizing high activity copper
the fate of the catalyst in the reactor, while in the methyl chromite catalysts. Among them (a) co-precipitation of

13
J Am Oil Chem Soc

Fig. 5  Dependence of a activity and b selectivity on surface area [79]

ammoniacal copper-chromium hydroxide, (b) co-impreg- However the heat-treated silica-free copper chromite cata-
nation of support with copper and chromium species, (c) lysts exhibit lower activity compared silica-based catalysts.
thermal decomposition of ammoniacal copper oxalate chro- These results indicate that the presence of silica not only
mate precursor are the most commonly used preparation improves acid resistance but also improves performance.
methods in the industry. Open literature information [83, 84] discussing the cata-
lyst degradation by acid attack and the information on the
Acid Resistance preparation of acid-resistant catalyst formulations is com-
paratively scanty. Ladebeck and Regula [83, 84] found
The presence of free fatty acid in the slurry process obvi- that copper chromite catalysts exhibit highest mechanical
ously makes great demands on the acid resistance of the stability and lowest copper solubility in feed/product. Alu-
catalyst. The copper-based catalysts are susceptible to acid mina supported copper catalysts show acid resistance close
attack as suggested in Eq. (13) and the copper metal can to that of copper chromite catalysts. According to these
elute as a soluble copper soap. In the presence of hydrogen, authors, the catalyst stability of the Cr-free copper catalysts
the copper soap can be reduced to Cu0 metal as shown in can be improved by avoiding high pressure in situ reduc-
Eq. (14) and deposit on the catalyst particle or reactor ves- tion or by applying an ex situ reduction technique so that
sel. Re-deposition of Cu metal on the catalyst can lead to water produced at the time of CuO reduction is removed
formation of larger Cu0 crystallite size. Thus, copper leach- from the catalyst pores before starting hydrogenation. In
ing not only impairs the catalyst performance due to Cu0 other words, the presence of water formed by reduction of
crystal growth (Fig. 7) but can also contaminate the fatty CuO can impair catalyst performance.
alcohol.
There are a few patents [73–76] teaching the ways Catalyst Filtration/Separation
to improve acid resistance by adding other components
including silica. Pohl et al. [73], Nierhaus et al. [74] and The slurry processes were designed and optimized by tak-
Schneider et al. [75, 76] claim the processes of making acid ing flow or hydrodynamic properties into consideration.
resistant copper chromite catalyst by incorporating silica Improvements in filtration rates were achieved by pre-
during precipitation step. Schneider et al. teach two ways cipitating uniform sized spherical particles [77, 78]. By
of improving acid tolerance. The first is by heat treatment optimizing the average particle size, Thakur et al. [77, 78]
[75] and the second [76] by incorporating acid resistance were able to qualify copper chromite catalysts for various
components such as silica and zirconia to copper chromite. filtration equipment used by fatty alcohol manufacturers.

13
J Am Oil Chem Soc

Fig. 6  Comparison of Cu–Cr catalyst preparation procedures: a activity, b life test, c filtration [80]

Fig. 7  Catalyst deactivation
due to copper crystallite growth
Well Dispersed Cu Particles Sintering of Cu° crystallite
More number of Cu active sites Fewer number of Cu active sites

Non‑Cr Copper Catalysts non-Cr-containing environmentally friendly catalysts. Vari-


ous researchers [85–101] focused their attention of devel-
The US Environmental Protection Agency, EPA and Euro- oping acid resistant highly active non-Cr catalysts based on
pean Union, REACH (Registration, Evaluation, Authoriza- CuZn, CuMn and CuFe compositions. Ladebeck and Reg-
tion and Restriction of Chemicals) regulations against the ula [83, 84] and Thakur et al. [87, 88] developed copper-
use of Cr(VI)-containing catalysts warranted a search for zinc-alumina catalyst systems that had equivalent or higher

13
J Am Oil Chem Soc

activity and acceptable acid tolerance. Schneider et al. [85, catalyst used in the fixed bed process are different from
86] and Chen [99, 100] focused their attention on copper- those employed for slurry phase catalysts. In addition to
manganese-alumina formulations for ester hydrogenation. meeting activity/selectivity targets, maldistribution of flow,
Research work reported by researchers at Utrecht Univer- temperature profile and pressure drop in the bed, and parti-
sity [90], Kao Soap [89, 97, 98, 101] and Roberts et al. [91, cle degradation also play critical roles in running a smooth
93] showed superior performance of CuFe supported on production campaign. Therefore, the catalyst loading war-
silica and alumina carriers for methyl ester hydrogenoly- rants special attention to avoid channeling through the bed
sis to fatty alcohols. Figure 8 compares the performance of and proper pre-activation ensures long catalyst life. The
various Cu catalysts with different promoters and also dem- catalyst chemist optimizes the mechanical integrity of the
onstrates the beneficial effect of Fe as a promoter [101]. catalyst particles so as to minimize particle breakup that
Van de Scheur and Staal [92], reported that the activity could cause pressure buildup inside the reactor or reduced
and selectivity for methyl acetate hydrogenation improved flow or flow channeling through the bed.
with increasing zinc content in CuZn systems. This increase
in activity was likely to be an effect that originates either Challenges for the Fixed Bed Catalyst Designer
from finely dispersed copper particles in close contact with
the zinc oxide phase or from the promotion of copper metal Activity, selectivity and the particle integrity are of utmost
crystallites by zinc oxide particles. A decrease in hydrocar- importance in determining the overall catalyst perfor-
bon formation with higher zinc content was attributed to mance. The catalyst formulator needs to choose a proper
the elimination of dehydrating surface sites (see Eq. 6). combination of active ingredients and promoters/modifiers
On the other hand, a joint collaborative effort between for high activity/selectivity, as well as a binder system for
Avantium and the University of Malaysia [102–106] physical integrity and chemical tolerance in acidic environ-
screened various Cu catalysts with different promot- ments. Higher activity per unit volume is manifested by
ers and different catalyst supports for hydrogenoly- higher throughputs and hence, translates into higher pro-
sis of fatty acid methyl esters at 300 °C and 300 bar. duction for the fatty alcohol manufacturer. Improved physi-
Their data showed the following trend for promoters cal integrity and attrition resistance minimize pressure drop
and supports respectively: Cu–Zn > Cu–Fe > Cu and problems. The catalyst is expected to perform under a wide
silica > zirconia > magnesia > alumina > titania. variety of conditions, such as liquid-phase or vapor-phase;
hydrogen pressures ranging from 30 to 300 bar, and with
Catalyst Development for Fixed Bed Catalyst different feeds, namely, wax or methyl ester derived from
palm or coconut.
While Table 1 provides the operating conditions, Fig. 4 It is well known that methyl or wax ester hydrogen-
illustrates the schematic of the fixed bed fatty alco- olysis is a pore diffusion controlled process. Thakur et al.
hol process. It is well recognized that the demand on the [6] demonstrated the advantages of a smaller and shaped
extrudate of 1.5 mm diameter as compared to the conven-
tional tablet technology of 3 or 4 mm diameter tablet in
terms of activity, selectivity and productivity. An increase
in void volume and higher surface area per unit volume in
the extrudates helps to decrease pressure drop problems
with the improvement of the performance over the tablets
as shown in Fig. 9. However, the shaped extrudates (tri-
lobe or star shaped) are traditionally weaker in mechanical
integrity as compared to the tablets because of their manu-
facturing process. Roberts et al. [107] developed a pro-
cess wherein an in situ calcium silicate (cement concrete)
is formed providing adequate mechanical strength to the
extrudates so that they can survive under operating condi-
tions and last for a long time in the reactor.
Extrudates were prepared by the process developed by
Roberts et al. [107] and Nebesh et al. [108] and evalu-
ated for high pressure fatty methyl ester hydrogenolysis at
250 bar, 180 °C and LHSV of 0.75 h−1. The product sam-
ples were analyzed for their saponification value as well
Fig. 8  Effect of different promoters on Cu catalysts [101] as by GC analysis. The saponification value in the liquid

13
J Am Oil Chem Soc

Saponification Value 50 Table 5  Comparison of trilobe extrudate versus tablet properties [6]


40 Trilobe extrudate catalyst Tablet catalyst
30
Diameter, inches 1/16 1/8
20
Bulk density, g/cc 0.7–0.9 1.3–1.5
10 % void volume 43 34
0 Surface area/volume 204 100
0 25 50 75 100 125 150 175 200
Hours-on-stream

Fig. 9  Activity comparison of extrudates vs tablets (triangles extru-


dates; squares tablets) [6] (saponification value = mg of KOH/g of
sample)

product sample decreases as the conversion increases. The


properties of the extrudates and tablets are given in Table 5.
While designing catalysts for the vapor-phase low pres-
sure process, which operates at higher temperatures than
the liquid-phase high pressure process (Table 1), Thakur
and Carrick [109] showed that the addition of alkali metal
to a copper chromite catalyst during extruding steps mini-
mizes undesired hydrocarbon formation and increases
selectivity.
Fig. 10  Effect of the addition of small amounts of lauric acid to a
feed containing methyl palmitate [112]. Conditions: (W/F) = 250 kg
catalyst (cat.) s/mol, C ester + acid (293 K, 1 atm) = 20 mmol/L,
Catalyst Deactivation
T = 197 °C

The natural oils used for making fatty esters contain 1 to


5 ppm of various catalyst poisons, which have been studied see the effect on the activity of these two catalysts. As the
by several authors [110–118]. free fatty acid concentration was increased in the feed to
Small amounts of free fatty acids (Van de Scheur et al. 0.05 and 0.14 mmol/L, the conversion dropped from 41 %
[112]), sulfur (Brands et al. [113], Twigg and Spencer [114], with no fatty acids to 34 and 16 %, respectively (Fig. 10).
Huang et al. [110]), chloride (Heldal and Mork [119], Twigg It is also found that the conversion of methyl palmitate is
and Spencer [114], Huang et al. [117]), water (Thakur et stable over time on stream in all the cases. Therefore, the
al. [111], Huang et al. [116]), phosphorous (Huang et al. detrimental effect of acid on the catalyst reactivity is due
[118]), and glycerine and monoglyceride (Thakur et al. to reaction inhibition rather than catalyst deactivation. The
[111]) present in the fatty acid methyl ester and wax ester original conversion level can be restored by switching from
feed may poison the catalyst and deactivate the reaction free acid spiked feed to pure methyl ester feed.
rate. The deactivation may be reversible (temporarily reduce
the activity) or irreversible (permanently reduce the activity) Effect of Sulfur on Catalyst Activity
depending on the nature of the impurity molecule.
Little is known about the nature of the sulfur compounds
Effect of Free Acid on Catalyst Activity present in fatty acid methyl esters prepared from natural
fats and oils. Possible candidates are degradation products
Small amounts of free fatty acids may be present in fatty of sulfur containing proteins (like thiols and sulfides) and
acid methyl ester and wax ester even after extensive puri- glucosinolates. The effect of sulfur poisoning for fatty ester
fication. Higher concentrations of free fatty acids present hydrogenolysis was reported by Brands et al. [113]. The
in the feed may lead to soap formation. The effect of free poisoning effect of sulfur on Cu based catalysts seems to be
fatty acids was studied by Van de Scheur et al. [112] in a permanent as it reacts with the copper site forming a strong
tubular flow reactor for methyl palmitate hydrogenolysis metal-sulfur bond, as indicated by the change in free energy
at 197 °C with 14 wt% Cu/9 wt% ZnO/SiO2 and 14 wt% formation (Twigg and Spencer [114], Brands et al. [113]):
Cu/SiO2 catalysts. Various concentrations of lauric acid
2Cu + H2 S → Cu2 S + H2 G0f = −122.1 kJ/mol (16)
(0.04–0.14 mmol/L) as free fatty acid spikes were used to

13
J Am Oil Chem Soc

Fig. 11  Sulfur deactivation mechanism adopted from Brands et al. [113]

The heat of adsorption of sulfur at the Cu crystal face nature of the sulfur containing molecules in FAME feed-
(−170 kJ/mol) is greater than the heat of formation of bulk stock for reliable estimation of catalyst life and reactor
copper sulfide (−140 kJ/mol) and the fact that the rate- design.
determining step of bulk copper sulfide formation is the Zn and Mn are excellent sulfur traps, and limit poison-
solid state diffusion of sulfur in copper or copper in sulfur- ing of copper catalysts by the removal of S-containing
implies that preferentially surface sulfides are formed. The compounds with the formation of zinc sulfide and manga-
difference in exothermicity between surface and bulk sulfi- nese sulfide respectively. This is attributed to the lower free
dation of copper catalysts is not very large, so concurrent energy of formation of bulk ZnS and MnS are −146 and
surface and bulk sulfidation is more likely. The mechanism −213 kJ/mol than that of Cu2S. A similar phenomenon was
proposed by Brands et al. [113] can be schematically pro- explained by Twigg and Spencer [114].
posed as in Fig. 11. Steps 1 and 2 are already discussed in
Zn + 1/2 S2 → ZnS �G0f = −146 kJ/mol (17)
the earlier section. In step 3, blocking of active sites and
thereby inhibiting the adsorption of the reactant molecules
MnO + H2 S → MnS + H2 O G0f = −213 kJ/mol (18)
occurs. This could be due to either adsorbed sulfur and/or
the remainder of the (organic) sulfur containing molecule. Brand et al. [113] showed that deactivation is faster for
In this step, sulfur may cause an increase in surface self- unpromoted Cu/SiO2 than Zn promoted Cu/SiO2 catalyst
diffusion (surface mobility), thus reducing catalytic activity in the presence of 356 ppm of 1-octadecanethiol in methyl
by sintering or restructuring of the active metal. palmitate hydrogenolysis carried out in a fixed bed reac-
Brands et al. [113] found that the type of sulfur com- tor at 197 °C and 80 bar pressure (Fig. 12). The octade-
pound has a strong influence on the rate of deactivation. canethiol breakthrough in outlet is also seen after 120 min
For example, octadecanethiol and dihexadecyl disulfide with Zn promoted Cu/SiO2 catalyst compared to 80 min
deactivate the catalyst at a slower rate than other types of with unpromoted Cu/SiO2 catalyst. Therefore, Zn acts as a
sulfur compounds (dihexadecyl sulfide, dibenzothiophene guard to the catalyst by preferentially adsorbing sulfur.
and methyl p-toluene sulfonate). In view of the differences Huang et al. [110] studied the poisoning effect of alkyl
in the deactivation behavior of the sulfur containing mol- thiols and dialkyl disulfides on the activity of dodecyl
ecules tested, Brands et al. [113] suggested establishing the methyl ester hydrogenation over co-precipitated Cu/Zn

13
J Am Oil Chem Soc

lower sulfur concentrations. These results suggest that the


disulfide species are more poisonous than the thiols and
can deactivate the catalyst at much lower S concentration
on the catalyst. Catalyst characterization by XRD EDS and
XPS measurements indicated that the sulfur species at low
concentrations (e.g. 0.2 mmol/gcatalyst ethanethiol), attack
the ZnO of the Cu/Zn catalyst forming chemically adsorbed
Zn–SR (where R = alkane). This interaction weakens the
synergistic effect between copper and zinc. At higher sul-
fur concentrations, both alkyl thiols (e.g. 15.1 mmol/gcatalyst
ethanethiol) and dialkyl disulfides also attack the surface
Cu of the Cu/Zn catalyst forming Cu7S4 or Cu31S16 as con-
firmed from XRD (Fig. 14).

Effect of Chlorides on Catalyst Activity


Fig. 12  Deactivation profile for methyl palmitate hydrogenolysis in
the presence of 356 ppm 1-octadecanethiol [113] (fixed bed reactor; The effect of chloride-containing species in the feed on the
200 mg catalyst; W/F = 2500 kgcat s mol−1 Cu and copper chromite catalysts were studied by Heldel
and Mork [119], Twigg and Spencer [114] and Huang et al.
[117]. Heldal and Mork [119] found that the amounts corre-
catalyst in a stirred tank reactor. Figure 13 compares the sponding to 25, 50 and 90 ppm Cl markedly decreased the
conversion of dodecyl methyl ester at various concentra- rate of hydrogenation with 39 %Cu, 32 % Cr and 2.5 % Mn
tion of sulfur species with different types of sulfur com- catalyst at 170 °C for hydrogenation of soybean oil. The
pounds. The ester conversion decreased rapidly when the presence of chloride species may exacerbate the poisoning
catalyst was exposed to ethanethiol, n-butanethiol and effect of other species (e.g. sulfur) as mentioned by Twigg
dodecanethiol. As the associated alkyl chain in the sulfide and Spencer [114]. They also noted that the deactivation
becomes longer, the deactivation of the catalyst becomes of the catalyst by chloride species may be due to blocking
faster. Depending on the types of alkyl thiol, the effects or modifying of the catalytic sites. The low melting point
on the activity of the Cu/Zn catalyst were different. It was of CuCl (430 °C) compared to Cu2S (1100 °C) gives them
found that the molecular toxicity increased continuously high surface mobility under operating conditions. This sug-
with the chain length, due to steric hindrance of alkyl gests that even extremely small amounts of copper halides
chains. Similar deactivation profiles were observed when are sufficient to provide mobile species that accelerate the
Cu/Zn catalyst was exposed dialkyl disulfides but at much sintering of Cu crystallites. Twigg and Spencer [114] also

Fig. 13  Conversion of dodecyl
methyl ester at different levels
of sulfur concentration in mmol/
gcatalyst and different types of
sulfur compounds [110]

13
J Am Oil Chem Soc

(c) 2Cu + O2− → Cu2 O + 2H+ + 2Cl− → 2CuCl + H2 O


(23)
The formation of ZnCl2 and CuCl at 15 mmol chloride
concentration/gcatalyst decreases the BET surface area from
19.1 to <0.3 m2/g and increases Cu (111) crystal size from
16.5 to >30 nm of the catalyst.

Effect of Water on Catalyst Activity

The reaction inhibition by water for fatty acid methyl ester


hydrogenolysis was found to be reversible causing tem-
porary deactivation [111]. The reaction was carried out
in a 1 L stirred reactor with a 400 g C12 methyl ester and
4–7.2 g copper–chromite catalyst at 207 bar and 280 °C.
The conversion of C12 fatty acid ester remained the same
Fig. 14  XRD Patterns of Cu/Zn catalysts after reaction with: (1) no up to 10 % water in the feed and is dropped by 4 % with
sulfur species; (2) 0.2 mmol/gcatalyst ethanethiol; (3) 5 mmol/gcatalyst 100 ppm water in the feed as shown in Fig. 15.
ethanethiol; (4) 15.1 mmol/gcatalyst ethanethiol; (5) 14.6 mmol/gcatalyst
The low activity of the Cu–Cr catalyst caused by
n-butanethiol; (6) 14.3 mmol/gcatalyst dodecanethiol [110]
water is mainly due to blocking of active sites by occlu-
sions. It not only occludes the active sites, but also may
pointed out that ZnO does not give any protection to Cu inhibit the reduction of cupric oxide to its zero valence
against chloride poisoning in contrast to the case of sulfide state active form. Fully reduced catalyst appears to be
poisoning, as the heat of formation of ZnCl2(s) (−121.8 kJ/ relatively immune to low levels of water contamination
mol) is much higher than that of CuCl(s) (−43.5 k/mol). (500 ppm). Thakur et al. [111] also found that the mech-
ZnCl2 has also low melting point (283 °C) and causes fur- anism by which soap (sodium laurate) deactivates the
ther poisoning and sintering of the catalyst. catalyst is similar to that suggested for water, i.e. active
Huang et al. [117] investigated the effect of the presence site occlusion and reduction inhibition (Fig. 14). Huang
of mono and dialkyl chlorides on Cu/Zn catalyst prepared et al. [116] had similar observations that the presence of
by co-precipitation method for hydrogenation of methyl water in the feed resulted in lower activity of the Cu–
laurate to dodecanol in a stirred tank reactor. The chloride Zn catalyst due to blocking the catalytic active sites by
concentration was varied from 0 to 0.5 mmol/gcatalyst. With the water occlusion of active sites. The concentration of
the increase of chloride concentration from 0 to 0.2 mmol/ water in the feed was increased from 0.1 to 2 wt%.and
gcatalyst, the conversion of methyl laurate decreased slightly. the conversion was decreased from 95 to 71 %. Fully
However, with a further increase in chloride concentration pre-reduced catalyst appears to be relatively immune
from 0.2 to 0.5 mmol/gcatalyst the conversion dropped from to low levels of water contamination, e.g. 1000 ppm of
90 to ~20 %. As expected, the poisoning effect of dichloro- water in methyl laurate. On the other hand, Tike and
alkane was more pronounced than monochloroalkane. They Mahajani [115] reported the enhanced hydrogenation
proposed the following mechanism, based on the analy- rate of palm stearin fatty acid for 5 % Ru/Al2O3 in the
sis of the species found both in liquid and on the catalyst, presence of water (2 wt%) compared to Ni based cata-
which is very similar to the other literature [114, 119]: lyst. Van der Scheur et al. [112] also reported no influ-
Scenario 1. Low chloride concentration: ence of added water on palmitate hydrogenolysis activ-
(a) Migration of Cl− from Cu to ZnO: ity of Cu/ZnO/SiO2 catalyst. They also commented on
different effects (positive, negative or neutral) of water
2Cl− + ZnO → ZnCl2 + O2− (19) on this kind of reaction, depending on the choice of the
2−
(b) Migration of O to Cu lattice: catalyst and the experimental conditions. Ladebeck and
Regula [85] found that the addition of large quantity
2Cu + O2− → Cu2 O (20) (1–2 %) of water to the ester and the subsequent treat-
Scenario 2: High chloride concentration: ment at 170 °C caused severe damage of the Cu/Al2O3
and Cu/ZnO/Al2O3 catalysts. This is mainly attributed to
(a) H+ + Cl− + CH3 OH → CH3 O+ H2 + Cl− (21) free fatty acid generated from the hydrolysis of methyl
(b) + −
2CH3 O H2 + 2Cl → ZnCl2 + 2H2 O + 2CH3 OH ester in the presence of water which leaches the metals
(22) from the catalyst.

13
J Am Oil Chem Soc

Fig. 16  Relative conversion of dodecyl methyl ester vs millimoles/


kg of permanent catalyst poisons [111]. φ2P-C2H4-Pφ2, bis(diphenyl
Fig. 15  Relative conversion of dodecyl methyl ester vs level of water phospheno)ethane; C18SO3Me, n-octadecyl methane sulfonate; C18Br,
and soap contamination [111] n-octadecyl bromide; C18Cl, n-octadecyl chloride; C18SH, n-octade-
cyl thiol

Effect of Phosphorous on Catalyst Activity

Huang et al. [118] reported the effect of phosphorous as a


contaminant in the fatty acid methyl ester feed. As little is
known about the type of phosphorous compounds present
in fatty acid methyl ester, they used trimethyl phosphate as
representative phosphorous compound for the model reac-
tion of hydrogenation of dodecyl methyl ester to dodecanol
with CuZn catalyst. When the concentration of trimethyl
phosphate was increased from 0 to 0.5 mmol/g of catalyst,
the conversion was found to decrease from 97 to 54 % and
the selectivity from 96 to 56 %. No catalytic activity was
observed at the concentration of 1 mmol/g of catalyst. The
rapid loss of catalyst activity may be attributed to the phys-
ical adsorption of trimethyl phosphate on the CuZn cata-
lyst and also to decreased BET surface area by occlusion
of active sites of the CuZn catalyst, as evidenced by XRD,
EDS, SEM, XPS and BET characterization.
Thakur et al. [111] presented the data of the perma- Fig. 17  Effect of glycerine and monoglyceride on C12 methyl ester
nent deactivation by irreversible adsorption of various hydrogenolysis [111]
feed impurities including phosphorous onto copper sites
using copper chromite as a catalyst for C12 methyl ester
hydrogenation as shown in Fig. 16 Interestingly, the effect decomposition products (propanol and 1,2-propanediol). It
of these poisons follows one trend, if we plot the relative also appears that the rate of temporary deactivation of the
activity versus the mole of poisons per g of catalyst. catalyst with increased concentration of glycerin, mono-
glyceride or 1,2-propane diol in the feed is the same. This
Effect of Glycerine on Catalyst Activity suggests that the molecular interaction with the copper
chromite surface may be the same for these molecules. In
Thakur et al. [111] also confirmed the completely revers- other words, these molecules may form a common interme-
ible or temporary deactivation of copper chromite cata- diate that interacts with the surface sites. In addition, their
lyst with glycerine and monoglyceride (Fig. 17) and their results show that the fresh oxide catalyst suffers higher

13
J Am Oil Chem Soc

activity loss when exposed to glycerine than the prereduced to fatty alcohol as reported by Fairwether et al. [124] and
catalyst. Acosta-Ramirez et al. [125]. The maximum yield of 91 %
of the desired alcohols were obtained with Milstein cata-
lyst for catalytic hydrogenation of coconut oil at 135 °C
Recent Developments on Catalysts for Fatty at 750 psig pressure [124]. Despite many advantages, this
Alcohol Chemistry process has long way to make industrial process as several
problems needs to be resolved before scale-up e.g., catalyst
Recent development on catalysts for fatty alcohol chemis- recovery, products separation, and disposal of toxic waste.
try is focused on searching for Cr free alternative formu-
lations both in fixed bed and slurry processes to reduce
the environmental pollution and to produce fatty alcohol Conclusions
under milder process conditions. Kandel et al. [120] stud-
ied Fe-induced promotional effect on Cu on mesoporous The literature pertaining to catalytic fatty alcohol processes
silica nanoparticles catalyst for hydrogenation of stearic was critically reviewed in this paper with special empha-
acid to 1-octadecanol in a batch reactor. The highest activ- sis on catalyst development for both slurry and fixed bed
ity (100 % conversion) with moderately high selectivity processes using fatty ester feeds. Challenges faced by the
(>95 %) was obtained with Cu/Fe mole ratio of 1. Increas- catalyst scientists while designing a catalyst for this com-
ing the relative loading of Fe in the catalyst leads to sig- plex chemistry are also discussed in detail. Environmen-
nificantly lower activity and selectivity. On the other hand, tally-friendly non-Cr Copper catalysts have been developed
increasing the loading of Cu leads to slightly lower activ- as an alternative to Cu–Cr catalyst, which are being used
ity (85 % conversion) but 100 % alcohol selectivity. Based commercially.
on this behavior, the authors have proposed a synergistic For the slurry process, a copper chromite catalyst pre-
hydrogenation mechanism where in situ-generated metal- pared by simultaneous precipitation at constant pH pro-
lic copper, which likely acts as a binding and activating duces a catalyst with narrow particle size distribution and
site for H2 and spills over to iron oxide creating an oxy- pore size distribution. These properties help to produce
gen vacancy. The carboxylic groups binds at the generated more activity and faster filterability. A CuCr catalyst is
oxygen vacancies and gets reduced by a second equivalent susceptible to acid attack. Addition of SiO2 to CuCr cata-
of hydrogen. A similar catalyst was developed by He et al. lyst improves the acid resistance and performance. For the
[121] for hydrogenation of ethyl stearate to stearyl alcohol fixed bed process, the new shape and design of the cata-
as a proof of concept carried out in an autoclave at 230 °C lyst was found to overcome the mass transfer limitation
and 30 bar pressure of hydrogen. The bimetallic catalyst thereby giving a higher performance at a lower tempera-
with a Cu/Fe mole ratio of 4/1 is more active than the cata- ture. In that case, the reactor design plays an important role
lyst with a Cu/Fe mole ratio of 1/4. They have attributed with efficient loading and minimized maldistribution of the
the higher activity of bimetallic catalyst with a Cu/Fe mole reactants.
ratio of 4/1 to the synergistic effects of Cu0/Cu+ species As the use of Cr(VI) is restricted due to stringent envi-
and Fe3O4, a large surface area and Lewis acid sites. ronmental regulations, recent focus has been to develop
Many reports in the literature [122–125] have given non-chrome catalyst, mainly from CuZn, CuMn and CuFe
importance recently to developing a homogenous catalyst groups. Very recent investigations have studied the interac-
for fatty acid methyl ester hydrogenation as it might have tion of Fe with Cu for higher activity. They have attributed
advantages in terms of selectivity and the ability to run at the higher activity of CuFe bimetallic catalyst to the syner-
milder process conditions. For example, Fairweather et al. gistic effects of Cu0/Cu+ species and Fe3O4, large surface
[124] obtained 98 % yield with 1.5 mol% five-coordinate area and Lewis acid sites.
Ru(II) PNN-pincer catalyst loading at 300 psig H2, 135 °C Catalyst deactivation studies point out that feed impuri-
temperature and 2 h reaction time. Mainly Ru- or Os-based ties such as P, S and Cl cause irreversible/permanent deac-
catalysts are used with different types of ligands connected tivation of Cu catalysts. On the other hand, the free acid,
to the metal (e.g. bidentate amino-phosphine, tetradentate triglyceride and water act as temporary poisons; with a sub-
imino-phosphine). Chakraborty et al. [122] developed Iron stantial recovery of activity upon switching to cleaner feed.
PNP-pincer complexes to avoid the precious metal to have Much understanding is needed on this aspect, especially
a more economical process. This catalyst showed compa- on the exploration on the kinds of contaminants (e.g. type
rable alcohol yield to ruthenium catalysts, but not yet as of sulfur, phosphorous and chloride species present in the
competitive as the most active ruthenium catalysts. The use feed).
of homogeneous catalysts might have another advantage in There have been some studies devoted to direct fatty
terms of application in direct hydrogenation of triglyceride acid hydrogenation to fatty alcohol with precious metal

13
J Am Oil Chem Soc

catalysts. The commercial applicability of this process is 19. Richardson A, Taylor J (1944) Process for forming alcohols or
limited as this process requires corrosion-resistant materi- esters. US Patent 2,340,343
20. Rabes IV, Schenck R (1948) Die Adkins’schen Kupfer-Chro-
als of construction and relatively expensive catalyst. mit-Katalysatoren und ihre Wirkungsursache. Zeitschrift für
Although there are a few researchers reporting advan- Elektrochemie und angewandte physikalische Chemie 52:37–39
tages of using supercritical solvents for fatty alcohol pro- 21. Reiner TW (1949) An improved laboratory preparation of cop-
duction, no plant has been built using this technology. The per-chromium oxide catalyst. J Am Chem Soc 71:1130
22. Stroupe JD (1949) An X-ray diffraction study of the copper
large solvent recycling and cumbersome product/solvent chromites and of the “copper-chromium oxide” catalyst. J Am
separation may be the extra hurdles in order to bring it to Chem Soc 71:569–572
the plant scale. 23. Adkins H, Burgoyne E, Schneider HJ (1950) The copper–
Few attempts have been made to develop a homogenous chromium oxide catalyst for hydrogenation. J Am Chem Soc
72:2626–2629
catalyst as this kind of catalyst can potentially exhibit very 24. De Nora V, de Bartholomaeis E (1956) Continuous hydrogena-
high activity with a small amount of catalyst and possess no tion of fatty material. US Patent 2,750,429
mass transfer limitation. The process scale-up constraint of 25. Toland W, Rafael S, Levine I (1957) Higher fatty alcohols. US
separating the homogenous catalyst from the liquid product Patent 2,776,323
26. Church J, Abdel-Gelil M (1957) Catalytic hydrogenation of
and the difficulty to make bulk quantities of this catalyst methyl laurate to lauryl alcohol. Ind Eng Chem 49:813–817
are the main obstacles for commercial viability. 27. Haidegger E, Karolyi J, Zalai A (1959) Die Erzeugung von Fet-
talkoholen durch katalytische Hochdruckhydrierung. I. Acta
Chim Acad Sci Hung 19:23
28. Karolyi J, Haidegger E, Hodossy L (1960) Erzeugung von fet-
References talkoholen mittels katalytischer hochdruckhydrierung. II. Acta
Chim Acad Sci Hung 20:157
1. Kenneally C (2001) Alcohols, higher aliphatic, survey. In: Kirk- 29. Pantulu AJ, Achaya KT (1964) Hydrogenolysis of saturated,
Othmer encyclopedia of chemical technology, vol 2. Wiley oleic and ricinoleic acids to the corresponding alcohols. J Am
2. Colin A (2013) Higher alcohols 2025. Houston Associates Oil Chem Soc 41:511–514
report 30. Eisenlohr K-H, Voeste T (1965) Process for the production of
3. Lo C, Cameron W (2012) Nexant ChemSystems PERP report fatty alcohols by catalytic hydrogenation of fatty acids and their
on “oleochemicals" derivatives. US Patent 3,180,898
4. Suyenty E, Sentosa H, Agustine M, Anwar S, Lie A, Sutanto E 31. Muttzall KMK (1966) High-pressure hydrogenation of fatty
(2007) Catalyst in basic oleochemicals. Bull Chem React Eng acid esters to fatty alcohols. Ph. D. thesis, University of Delft,
Catal 2:22–31 The Netherlands
5. Behr A, Westfechtel A, Gomes JP (2008) Catalytic processes 32. Muttzall KMK, van der Berg PJ (1971) Hydrogenation of fatty
for the technical use of natural fats and oils. Chem Eng Technol acid esters to fatty alcohols. In: Proceedings of the European
31:700–714 symposium chemical reaction engineering (1968), 4th edn. Per-
6. Thakur D, Okonek D, Hoelzle M (2008) Paper presented at gamon Press Ltd., Oxford, pp 277–285
PRO 3/EXH 2: processing exhibitor session; the 99th AOCS 33. Frazee JR, Martin B, Brundrett C (1973) Supported copper
annual meeting and expo chromite catalyst. US Patent 3,756,964
7. Noweck K, Grafahrend W (2012) Fatty alcohols. Ullmann’s 34. Montgomery S (1973) Process for producing copper chromite
encyclopedia of industrial chemistry, vol 14. Wiley-VCH Verlag catalysts. US Patent 3,767,595
GmbH & Co. KGaA, Weinheim, pp 117–141 35. Boerma H (1976) In: Delmon B, Jacobs P, Poncelet G (eds)
8. Hansley V.L. (1937) Reduction of esters. US Patent 2,096,036 Preparation of copper and zinc chromium oxide catalysts for the
9. Hansley V.L (1939) Ester condensations. US Patent 2,158,071 reduction of fatty esters to alcohols. In: International Sympo-
10. Kastens M, Peddicord H (1949) Alcohols by sodium reduction. sium on preparation of catalysis, vol 1. Elsevier, pp 105-118
Ind Eng Chem 41:438–446 36. Monick JA (1979) Fatty alcohols. J Am Oil Chem Soc
11. Adkins H, Folkers K (1931) The catalytic hydrogenation of 56:853A–860A
esters to alcohols. J Am Chem Soc 53:1095–1097 37. Voeste T, Schmidt HJ, Marschner F (1981) Continuous process
12. Norman W (1931) Über die katalytische Reduktion der Carbox- of producing fatty alcohols. US Patent 4,259,536
ylgruppe. Angew Chem 44:714–717 38. Buchold H (1983) Natural fats and oils route to fatty alcohols.
13. Schrauth W, Schenk O, Stickdorn K (1931) Über die Her- Chem Eng 90:42–43
stellung von Kohlenwasserstoffen und Alkoholen durch 39. Voeste T, Buchold H (1984) Production of fatty alcohols from
Hochdruck-Reduktion von Fettstoffen. Ber Dtsch Chem Ges fatty acids. J Am Oil Chem Soc 61:350–352
64:1314–1318 40. Kreutzer UR (1984) Manufacture of fatty alcohols based on
14. Schmidt O (1931) Die Katalytische Hydrierung der Carbox- natural fats and oils. J Am Oil Chem Soc 61:343–348
ylgruppe in organischen Verbindungen, Insbesondere solchen 41. Schuett H (1989) Process for obtaining fatty alcohols from free
höheren Molekulargewichts. Ber Dtsch Chem Ges 64:2051–2053 fatty acids. US Patent 4,804,790
15. Adkins H, Folker K, Conner R (1937) Method of hydrogenating 42. Fleckenstein T (1991) Process for the hydrogenation of fatty
esters. US Patent 2,091,800 acid methyl ester mixtures. US Patent 5,043,485
16. Adkins H (1940) Hydrogenation role of the catalyst. Ind Eng 43. Demmering G, Heck S, Friesenhagen L (1994) Process for the
Chem 32:1189–1192 production of fatty alcohols. US Patent 5,364,986
17. Rittmeister W (1941) Process for the reduction of fatty acids to 44. Poels EK, Polman DRE, Vresswijk JJ (1992) Alcohol produc-
alcohols. US Patent 2,248,465A tion by the hydrogenation of fatty acid esters. GB 2,250,287 A1
18. Rittmeister W (1959) Process for the production of unsaturated 45. Wilmott M, Harrison G, Scarlett J, Wood M, McKinley D
fatty alcohols. US Patent 3,193,586A (1992) Fatty alcohols. US Patent 5,138,106

13
J Am Oil Chem Soc

46. Wilmott M, Harrison G, Scarlett J, Wood M, McKinley D 67. Miyake T, Makino T, Taniguchi S-I, Watanuki H, Niki T,
(1992) Process for the production of fatty alcohols. US Patent Shimizu S, Kojima Y, Sano M (2009) Alcohol synthesis by
5,157,168 hydrogenation of fatty acid methyl esters on supported Ru–Sn
47. Buchold H, Gartner F-J, Mallok G, Schlichting E, Stonner H-M and Rh–Sn catalysts. Appl Catal A 364:108–112
(1997) Process for preparing wax esters and hydrogenation of 68. Echeverri D, Marin J, Restrepo G, Rios L (2009) Characteri-
wax esters to fatty alcohols. US Patent 5,608,122 zation and carbonylic hydrogenation of methyl oleate over Ru–
48. Boensch R, Noweck K (2013) Method for the production of Sn/Al2O3: effects of metal precursor and chlorine removal. Appl
fatty alcohols. US Patent 8,426,654 Catal A 366:342–347
49. Appleton P, Wood M, Wild R (2014) Process for producing 69. Zhang L, Zheng X, Li R (2015) Effect of Lanthanum on catalytic
fatty alcohols from fatty acids. US Patent 8884078 performance of Ru/Al2O3 catalyst for hydrogenation of esters to
50. van de Scheur FT, Vreeswijk JJ (1994) Process for the produc- corresponding alcohols. Chem Res Chin Univ 31:615–620
tion of alcohols. WO 1994006738 A1 70. Rozmysłowicz B, Kirilin A, Aho A, Manyar H, Hardacre C,
51. van den Hark S, Härröd M, Møller P (1999) Hydrogenation of Warna J, Salmi T, Murzin D (2015) Selective hydrogenation of
fatty acid methyl esters to fatty alcohols at supercritical condi- fatty acids to alcohols over highly dispersed ReOx/TiO2 cata-
tions. J Am Oil Chem Soc 76:1363–1370 lyst. J Catal 328:197–207
52. van den Hark S, Harrod M (2001) Hydrogenation of oleochemi- 71. Miya B, Sawamoto Y, Hashiba K, Hisamitsu H (1979) Process
cals at supercritical single-phase conditions: influence of hydro- for preparation for copper-iron-aluminium catalysts and cata-
gen and substrate concentrations on the process. Appl Catal lysts prepared by the process. US Patent 4,144,198
210:207–215 72. Miya B, Miya F, Nawashiro Y (1981) Process for preparation
53. Brands DS, Poels EK, Dimian AC, Bliek A (2002) Solvent- of copper-iron-aluminium hydrogenation catalyst. US Patent
based fatty alcohol synthesis using supercritical butane: thermo- 4,278,567
dynamic analysis. J Am Oil Chem Soc 79:75–83 73. Pohl J, Carduck F-J, Goebel G (1989) Acid-resistant catalysts
54. Brands DS, Pontzen K, Poels E, Dimian A, Bliek A (2002) Solvent- for the direct hydrogenation of fatty acids to fatty alcohols. US
based fatty alcohol synthesis using supercritical butane: flowsheet Patent 4,855,273
analysis and process design. J Am Oil Chem Soc 79:85–91 74. Nierhaus W, Pohl J, Goebel G (1992) Copper/manganese cata-
55. Zhilong Y (2011) Biodiesel quality, emissions and by-products. lysts. WO 1992/04119
In: Montero G (ed) Research on hydrogenation of FAME to 75. Schneider M, Maletz G, Kochloeft K (1993) SiO2-containing
fatty alcohols at supercritical conditions, Ch. 11. INTECH, New copper oxide-chromium oxide catalyst for the hydrogenation of
York, pp 170–180 fatty acids and fatty esters. US Patent 5,217,937
56. Liang S, Liu H, Wang W, Jiang T, Zhang Z, Han B (2012) 76. Schneider M, Maletz G, Kochloeft K (1993) Acid-resistant cop-
Hydrogenation of methyl laurate to produce lauryl alcohol per oxide-chromium oxide catalyst for the hydrogenation of
over Cu/ZnO/Al2O3 with methanol as the solvent and hydrogen fatty acids and fatty esters. US Patent 5,206,203
source. Pure Appl Chem 84:779–788 77. Thakur D, Palka E, Sullivan T, Nebesh E, Roberts B (1992)
57. Broadbent H, Campbell G, Bartley W, Johnson J (1959) Rhe- Process for preparing catalyst with copper or zinc and with
nium and its compounds as hydrogenation catalysts. III. Rhe- chromium, molybdenum, tungsten, or vanadium, and product
nium Heptoxide. J Org Chem 24:1847–1854 thereof. US Patent 5,134,108
58. Trivedi BC (1978) Catalytic hydrogenation of fatty acids. US 78. Thakur D, Palka E, Sullivan T, Nebesh E, Roberts B (2000)
Patent 4,104,478 Hydrogenation catalyst, process for preparing and process of
59. Trivedi BC, Grote D, Mason ThO (1981) Hydrogenation of using said catalyst. US Patent 6,054,627
carboxylic acids and synergistic catalysts. J Am Oil Chem Soc 79. Rieke RD, Thakur DS, Roberts BD, White GT (1997) Fatty
58:17–20 methyl ester hydrogenation to fatty alcohol part I: correlation
60. Narasimhan CS, Deshpande VM, Ramnarayan K (1989) Selec- between catalyst properties and activity/selectivity. J Am Oil
tive hydrogenation of methyl oleate to oleyl alcohol on mixed Chem Soc 74:333–339
ruthenium-tin boride catalysts. Appl Catal 48:L1–L5 80. Rieke R, Thakur DS, Roberts BD, White GT (1997) Fatty
61. Yoshino K, Kajiwara Y, Takaishi N, Inamoto Y, Tsuji J (1990) methyl ester hydrogenation to fatty alcohol part II: process
Hydrogenation of carboxylic acids by rhenium–osmium bimet- issues. J Am Oil Chem Soc 74:341–345
talic catalyst. J Am Oil Chem Soc 67:21–24 81. Prasad R (1993) A new precursor for the preparation of novel
62. Nagahara E, Itoi Y (1995) Catalyst for direct reduction of copper chromite catalysts. Stud Surf Sci Catal 75:1747–1750
carboxylic acid, process for preparation there of and pro- 82. Prasad R, Singh P (2011) Applications and preparation methods
cess for preparation of alcohol using the catalyst. US Patent of copper chromite catalysts: a review. Bull Chem React Eng
5,426,246 Catal 6:63–113
63. Tahara K, Nagahara E, Itoi Y, Nishiyama S, Tsuruya S, Masai 83. Ladebeck J, Regula T (1999) Fatty methyl ester hydrogena-
M (1997) Liquid-phase hydrogenation of carboxylic acid on tion: application of chrome free catalysts. Stud Surf Sci Catal
supported bimetallic Ru–Sn–alumina catalysts. Appl Catal A 121:215–220
154:75–86 84. Ladebeck J, Regula T (2001) Catalysis of organic reactions. In:
64. Pouilloux Y, Autin F, Guimon C, Barrault J (1998) Hydrogena- Ford ME (ed) Copper-based chromium free hydrogenation cata-
tion of fatty esters over ruthenium-tin catalysts; characterization lysts. Marcel Dekker (Publ), New York, pp 403–413
and identification of active centers. J Catal 176:215–224 85. Schneider M, Kochloeft K, Maletz G (1995) Catalyst and pro-
65. Pouilloux Y, Autin F, Piccirilli A, Guimon C, Barrault J (1998) cess for hydrogenation of carboxylic acid alkyl esters to higher
Preparation of oleyl alcohol from the hydrogenation of methyl alcohols. US Patent 5,386,066
oleate in the presence of cobalt-tin catalysts. Appl Catal A 86. Schneider M, Kochloeft K, Maletz G (1995) Chromium-free
16:65–75 catalyst for the hydrogenation of organic compounds. US Patent
66. Toba M, Tanaka S, Niwa S, Mizukami F, Koppany Z, Guczi L, 5,403,962
Cheah K-Y, Tang T-S (1999) Synthesis of alcohols and diols 87. Thakur D, Roberts B, Sullivan T, Vlacheck A (1992) Hydrogen-
by hydrogenation of carboxylic acids and esters over Ru–Sn– ation catalyst, process for preparing and process of using said
Al2O3 catalysts. Appl Catal A 189:243–250 catalyst. US Patent 5,155,086

13
J Am Oil Chem Soc

88. Thakur D, Roberts B, Sullivan T, Vlacheck A (1994) Hydrogen- 108. Nebesh E, Kelly DG, Novak LT (1992) Copper chromite
ation catalyst, process for preparing and process of using said catalyst and process for preparation said catalyst. US Patent
catalyst. US Patent 5, 345,005 5,124,295
89. Matsuda M, Horio M, Tsukada K (1992) Process for producing 109. Thakur DS, Carrick WJ (2015) Copper chromite hydrogenation
hydrogenation catalyst. US patent 5,120,700 catalysts for production of fatty alcohols. US Patent 9,120,086
90. van Beijnum J, van Dillen AJ, Geus J (1993) Hydrogenolysis 110. Huang H, Cao G, Wang S (2014) An evaluation of alkylthi-
reaction and catalyst suitable therefor. US Patent 5,198,512 A ols and dialkyl disulfides on deactivation of Cu/Zn catalyst in
91. Roberts BD, Thakur DS, Sullivan TJ, Plundo RA (1993) Hydro- hydrogenation of dodecyl methyl ester to dodecanol. J Ind Eng
genation catalyst, process for preparing and process for using Chem 20:988–993
said catalyst. US Patent 5,243,095 111. Thakur DS, Roberts BD, White GT, Rieke RD (1999) Fatty
92. van de Scheur F, Th Staal L H (1994) Effects of zinc addition methyl ester hydrogenation to fatty alcohol: reaction inhibi-
to silica supported copper catalysts for the hydrogenolysis of tion by glycerine and monoglyceride. J Am Oil Chem Soc
esters. Appl Catal 108:63–83 76:995–1000
93. Roberts BD, Thakur DS, Sullivan TJ, Plundo RA (1995) Hydro- 112. van de Scheur FT, Sai GU-A, Bliek A, Staal L (1995) The effect
genation catalyst and process for preparing same. US Patent of free fatty acid on the reactivity of copper-based catalysts for
5,418,201 the hydrogenolysis of fatty acid methyl esters. J Am Oil Chem
94. Brands DS, Poels EK, Bliek A (1996) The relation between pre- Soc 72:1027–1031
treatment of promoted copper catalysts and their activity in hydro- 113. Brands DS, Sai GU-A, Poels EK, Bliek A (1999) Sulfur
genation reactions. Stud Surf Sci Catal 101:1085–1094 deactivation of fatty ester hydrogenolysis catalysts. J Catal
95. Chen Y, Chang C (1997) Cu-B2O3/SiO2, an effective catalyst 186:169–180
for synthesis of fatty alcohol from hydrogenolysis of fatty acid 114. Twigg MV, Spencer MS (2001) Deactivation of supported cop-
esters. Catal Lett 48:101–104 per metal catalysts for hydrogenation reactions. Appl Catal A
96. Brands DS (1998) The hydrogenolysis of esters to alcohols over 212:161–174
copper containing catalysts. Ph. D. thesis, University of Amster- 115. Tike MA, Mahajani VV (2007) Kinetics of hydrogenation of
dam, The Netherlands palm stearin fatty acid over Ru/Al2O3 catalyst in presence of
97. Tsukada K, Hattori Y, Mimura T (1996) Method for preparing small quantity of water. Ind J Chem Technol 14:52–63
copper-containing hydrogenation reaction catalyst and method 116. Huang H, Cao G, Fan C, Wang S, Wang S (2009) Effect of
for producing alcohol. US Patent 5,554,574 water on Cu/Zn catalyst for hydrogenation of fatty methyl ester
98. Matsuda M, Horio M, Tsukada K, Sotoya K, Abe H, Tsushima to fatty alcohol. Korean J Chem Eng 26:1574–1579
R (1994) The production of fatty alcohols and their amino 117. Huang H, Wang S, Wang S, Cao G (2010) Deactivation mech-
derivatives from Coco fatty acid methyl esters In: Applewhite anism of Cu/Zn catalyst poisoned by organic chlorides in
T (ed) Proceedings of the World conference on lauric oils: hydrogenation of fatty methyl ester to fatty alcohol. Catal Lett
sources, processing, and applications, chapter 7, pp 64–71 134:351–357
99. Chen J. P. (2005) Preparation and use of non-chrome catalysts 118. Huang H, Cao G, Wang S (2013) An evaluation of trimethyl
for Cu/Cr catalyst applications. US Patent 6,916,457 phosphate on deactivation of Cu/Zn catalyst in hydrogenation
100. Chen JP (2014) Copper catalyst for dehydrogenation applica- of dodecyl methyl ester. Korean J Chem Eng 30:1710–1715
tion. US Patent 8,828,903 119. Haldal JA, Mork P (1982) Chlorine-containing compounds as
101. Hattori Y, Yamamoto K, Kaita J, Matsuda M, Yamada S (2000) copper catalyst poisons. J Am Oil Chem Soc 59:396–398
The development of nonchromium catalyst for fatty alcohol 120. Kandel K, Chaudhary U, Nelson NC, Slowing II (2015) Syner-
production. J Am Oil Soc 77:1283–1288 gistic interaction between oxides of copper and iron for produc-
102. Zuzaniuk V, Gruter G-J, Veringa M, Sijpkes A, van der Puil N, tion of fatty alcohols from fatty acids. ACS Catal 5:6719–6723
De Keijzer J, Hamid SBA (2007) Catalyst and process develop- 121. He L, Li X, Lin W, Li W, Cheng H, Yu Y, Fujita S-I, Arai M,
ment for production of oleochemicals from palm oil using high Zhao F (2014) The selective hydrogenation of ethyl stearate to
throughput experimentation. North American Catalysis Society stearyl alcohol over Cu/Fe bimetallic catalysts. J Mol Catal A:
meeting Chem 392:143–149
103. Spikes A, van der Puil N, van der Brink P-J, de Keijzer AHJF, 122. Chakraborty S, Dai H, Bhattacharya P, Fairweather NT, Gib-
Hamid SBA (2005) Chromium-free catalysts of metallic Cu and at son MS, Krause JA, Guan H (2014) Iron-based catalysts
least one second metal. WO 2005070537 A1 for the hydrogenation of esters to alcohols. J Am Chem Soc
104. bin Ma’amor A, Hamid SBA (2006) Combinatorial technology 136:7869–7872
in heterogenous copper-based catalysts. In: Proceedings of the 123. Tan X, Wang Y, Liu Y, Wang F, Shi L, Lee K-H, Lin Z, Lv H,
1st international conference on natural resources engineering Zhang X (2015) Highly efficient tetradentate ruthenium catalyst
and technology; Putrajaya, Malaysia, pp 255–261 for ester reduction: especially for hydrogenation of fatty acid
105. Spikes A, van der Puil N, van der Brink P-J, Hamid SBA, de esters. Org Lett 17:454–457
Keijzer AHJF (2007) Chromium-free catalysts of metallic Cu 124. Fairweather NT, Gibson MS, Guan H (2014) Homogeneous
and at least one second metal. US 2007/0207921 A1 hydrogenation of fatty acid methyl esters and natural oils under
106. Spikes A, van der Puil N, van der Brink P-J, Hamid SBA, neat conditions. Organometallics 34:335–339
de Keijzer AHJF (2010) Chromium-free catalysts of metal- 125. Acosta-Ramirez A, Bertoli M, Gusev DM, Schlaf M (2012)
lic Cu and at least one second metal. US Patent Application Homogenous catalytic hydrogenation of long-chain esters by
2010/0087312 A1 an osmium pincer complex and its potential application in the
107. Roberts BD, Carrick WJ, Thakur DS (1999) Shaped hydrogena- direct conversion of triglycerides into fatty alcohols. Green
tion catalyst and processes for their preparation and use. US Chem 4:1178–1188
Patent 5,977,010

13

You might also like