Vous êtes sur la page 1sur 16

Applied Catalysis A, General 543 (2017) 82–97

Contents lists available at ScienceDirect

Applied Catalysis A, General


journal homepage: www.elsevier.com/locate/apcata

Feature Article

Applications of light olefin oligomerization to the production of fuels and MARK


chemicals
Christopher P. Nicholas
Exploratory Catalysis and Materials Research, Honeywell UOP, 25 East Algonquin Road, Des Plaines, IL 60017, USA

A R T I C L E I N F O A B S T R A C T

Keywords: The oligomerization of the light olefins ethene, propene, and butenes into fuels and chemicals has been in-
Catalysis vestigated and commercially practiced for many years. While the area appears on the surface to be mature, many
Zeolite advances have been made in recent years. In this feature article, I discuss the mechanisms of reaction and
Nickel showcase catalysts and processes useful for oligomerization from both the open and patent literature.
Resin
Commercially practiced processes are spotlighted. Among the catalysts utilized in the art are acidic catalysts
Solid phosphoric acid
such as solid phosphoric acid and zeolites, as well as metal based catalysts including aluminum alkyls, and nickel
Metallacycle
Fuels and zirconium based complexes and solids. A short section on catalysts and processes which utilize a metalla-
Chemicals cycle mechanism in order to achieve high selectivity to 1-hexene or 1-octene is followed by a discussion of multi-
functional materials possessing both acid and metal active sites. Finally, processes where oligomerization is a key
step in a multi-step or multi-reaction process are discussed.

1. Introduction is divided according to the mechanism utilized by the catalysts, with


acid-catalyzed carbenium chemistry and metal catalyzed 1,2-insertion
The oligomerization of the light olefins ethene, propene, and bu- the two most frequently utilized mechanistic routes for oligomerization.
tenes into fuels and chemicals has been investigated and commercially Both types of catalysts are used in currently commercially practiced
practiced for many years [1]. Initially, non-catalytic thermal methods processes. In these reactions, the relative rates of oligomerization and
had been utilized to convert light olefins into more useful compounds side reactions such as isomerization (both double bond and skeletal),
[2–4], but Ipatieff’s discovery of solid phosphoric acid catalyzed oli- cracking (mono- and bi-molecular), aromatization, hydride transfer,
gomerization 80 years ago sparked the first widespread use of oligo- and coking are a strong function of structure and reaction conditions
merization as a method of synthesizing transportation fuels [5,6]. Mobil (e.g. temperature, pressure, contact time) [13,14]. In general, cracking
then advanced acidic oligomerization technology with commercial reactions and hydride transfer reactions often leading to aromatics and
application of MFI zeolite in the conversion of light olefins derived from coke precursors are favored at higher temperatures (> 300 °C) [15],
methanol to fuels in their New Zealand refinery from 1985 to 1997. A with oligomerization favored at lower temperature (< 200 °C) [16].
key step in the overall process is the oligomerization of light olefins to Thus, the selectivity to specific products is dependent both on catalyst
gasoline and distillate [7]. In the meantime, Ziegler had discovered and process conditions.
metal catalyzed oligo/polymerization [8], several of which catalyst Understanding the impact of mechanism on oligomer molecular
types have since been commercialized. While thermal methods have weight distribution is also important. Two are frequently observed,
been recently investigated again, and radical mediated oligomerization both of which describe asymmetric oligomer distributions. Because zero
is also feasible, catalytic methods dominate the market due to the insertions is the monomer, negative insertions by definition cannot
control afforded over the product [9,10]. occur, but production of any length chain is feasible, an asymmetric
In this feature article, I discuss recent developments in catalytic distribution of oligomers is mathematically necessary.
olefin oligomerization of C2eC5 olefins, with highlights both from the A long known mathematical function, the Poisson distribution [17],
patent and open literature that have occurred, largely since O’Connor is the less frequently observed and can be utilized to describe dis-
and Kojima’s review [11]. To help focus the text, alcohol and other tributions of oligomers where production of each oligomer number is
oxygenated molecules are not covered as feedstocks here unless a independent of others [18]. Acid catalysts need to build a high con-
specific step to produce olefins has been reported, although biomass centration of dimers before trimers are produced, therefore, acid cata-
conversion via oligomerization routes has been successful [12]. The text lysts should never give a Poisson distribution of oligomers since this key

E-mail address: Christopher.Nicholas@honeywell.com.

http://dx.doi.org/10.1016/j.apcata.2017.06.011
Received 17 March 2017; Received in revised form 5 June 2017; Accepted 7 June 2017
Available online 09 June 2017
0926-860X/ © 2017 Elsevier B.V. All rights reserved.
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

condition is not met. Metal based catalysts sometimes give Poisson


distributions, but chain termination needs to be independent of
monomer insertion as well as chain length independent, and oligomer
reinsertion cannot occur.
The Schultz-Flory distribution can be used to describe oligomer
distributions where chain termination is kinetically competitive with
the rate of propagation and/or reinsertion of oligomers occurs [19,20].
Cracking or other side reactions should not be present at kinetically
relevant rates. Most metal catalysts will therefore give Schultz-Flory
distributions of oligomers; these tend to be broader than Poisson dis-
tributions [21].
Following a discussion of catalysts using acid and insertion me-
chanisms, I then explore the typically homogeneous catalysts which
utilize a metallacycle mechanism to oligomerize ethylene and other
light olefins into dimers and trimers. These catalysts are the base of a
growing industry to produce C4eC8 olefins as comonomers for poly-
ethylene and polypropylene production. After discussing bifunctional
catalysts which utilize both metal and acid sites/mechanisms, I finish
by looking at several combination processes where oligomerization is Fig. 1. Carbenium ion mechanism utilized in acid catalyzed oligomerization reactions.
one of the key steps. The first olefin inserted is black, the second blue, and the third red. Numeration is de-
scribed in the text. (For interpretation of the references to colour in this figure legend, the
2. Acid catalyzed oligomerization reader is referred to the web version of this article.)

While olefin oligomerization is a deactivation or undesired reaction another olefin in reaction 2 to give a longer carbon chain. The inser-
mechanism in many acid catalyzed transformations of olefins [22], it is tion/chain growth can continue (reaction 3) or terminate (reaction 4) to
also a useful method of carbon–carbon bond formation, provided that yield an olefin and the starting acid site. With substitution on the olefin,
proper catalyst and conditions are utilized. For over 80 years, C2eC5 each olefin addition (dimer, trimer, etc.) yields a branch in the hy-
light olefins have been oligomerized over acidic catalysts to fuels, drocarbon chain. Isomerization is also quite prevalent in many systems
principally gasoline, but recently to distillates as well. Market sizes are with H-shift (reaction 5) and alkyl shifts (reaction 6) frequently oc-
difficult to estimate as the fuels produced via oligomerization are ty- curring to yield more thermodynamically stable carbenium ions. Thus,
pically not accounted separately from other refinery sources, but an true oligomerization products are often not the only products formed.
installed base of approximately 60 oligomerization units primarily Reaction 1 can be responsible for double bond migration and E to Z
converting propene, butenes, or combinations thereof with nameplate transitions. Once the planar carbenium ion is formed, if reversed, the
capacity of about 67.5 M barrels/year of feed remain. Just under 650 position of the olefin and relative E or Z configuration can change from
oil refineries exist in the world, so 9% of refineries currently practice the initial olefin. The particular acidic solid utilized and confinement
oligomerization [23]. effects drive differentials in the relative rates of the multiple reactions
In addition to fuels, this total capacity includes the primary pro- as Iglesia and coworkers have shown [32].
duction of olefin oligomers as precursors to petrochemicals including
branched alkylbenzenes (BAB), alkylated phenols, and oxo alcohols. 2.1. Solid phosphoric acid (SPA)
Marketshare of BAB for sulfonation and use as a detergent has been
significantly declining in favor of the more readily biodegradable linear Phosphoric acid, as a high concentration aqueous solution, had been
alkylbenzenes (LAB, Section 3) [24], but still sees use as oilfield che- shown to be catalytically active for the oligomerization of olefins
micals. The largest alkylated phenol use is in nonylphenolethoxylate [33,34], with significantly higher yields than thermal polymerization
non-ionic surfactants, where over 300 million pounds are produced and with activity dependent on the concentration of H3PO4 [35,36]. It
yearly, though market share will decline rapidly in upcoming years wasn’t until the development of solid phosphoric acid (SPA) that in-
[25,26]. dustrial processes for olefin oligomerization were commercialized
Oxo alcohols synthesized from mixed octenes and nonenes see end [37,38]. 80 years after Ipatieff’s discovery [5], SPA continues to be
uses in plasticizers for polyvinylchloride (PVC) including diisono- utilized commercially due to its high selectivity to gasoline and cost
nylphthalate (DINP), the hydrogenated non-phthalate version cyclo- effectiveness.
hexane dicarboxylic acid diisononyl diester (DINCH®), diisode- In addition to UOP’s Catalytic Condensation™ process [39], of which
cylphthalate (DIDP), trimellitates, and adipates. The use of DINP and more than 250 units have been licensed since 1935 (also called Cat-
DINCH® in particular have been increasing significantly as di-2-ethyl- Poly) [40], and Standard Oil’s development of a similar process [41],
hexylphthalate (DEHP) has declined due to environmental concerns. Of Sasol have been practicing a SPA based oligomerization process [42]
the approximately 6 million ton/yr plasticizer market, these molecules and have published much research in the area. SPA is formed by
had about 33% market share in 2014, with the total share expected to combining H3PO4 with silica sources such as the natural products kie-
grow in the next several years [27]. Other significant uses for oxo al- selguhr or diatomite [43], synthetic silicas [44,45], or quartz followed
cohols are in alkylethoxylate non-ionic surfactants whose market share by extrusion [46]. High silica content solids are traditionally utilized,
is also increasing [28]. but even low silica content kieselguhr with CaO and MgO contents
Multiple acidic solids [29] have been utilized as oligomerization above 3 wt% (combined) are acceptable as long as the bulk density is
catalysts, ranging from the classical silica-supported phosphoric acid, to less than about 0.3 g/mL [47]. During the reaction between H3PO4 and
acidic ion-exchange resins, amorphous silica-aluminas, zeolites and the SiO2 support, a number of silicon phosphates and hydrogen phos-
even acidic clays and sulfated metal oxides [30]. The primary me- phates including orthophosphate Si5O(PO4)6, pyrophosphate SiP2O7,
chanism utilized over these materials is a classic carbenium route [31], hydrogen phosphate Si(HPO4)2·H2O, and tripolyphosphate SiHP3O10
shown in Fig. 1. are formed [48]. The orthophosphate is the preferred phase for activity
The Brønsted acid site attacks the olefin in reaction 1 leading to the [49], the pyrophosphate for strength [46,50] and the hydrogenpho-
formation of an ion-pair. The carbenium ion formed can react with sphate a non-preferred phase. As for many of the other catalyst systems

83
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Fig. 2. The phosphate ester mechanism exhibited by SPA. Reprinted with permission
from Ind. Eng. Chem. Res. 45 (2006) 578–84. Copyright 2006 American Chemical
Society.

utilized for olefin oligomerization, it may be beneficial to have large


macropores [51].
The acidity (and activity) of SPA has been shown to be due to
phosphoric acid oligomers supported on the silicon phosphate phases.
Here, the mechanism involves not a true acid derived carbocation, but
instead a phosphate alkylester with significant interaction between the
cationic protonated olefin and anionic phosphate, as shown in Fig. 2
[48,52]. This mechanism effectively controls the product selectivity to Fig. 3. Use of multi-dimensional NMR to show SPA product speciation differences from
gasoline, but selectivity can be adjusted towards distillate by decreasing MTW zeolite based catalysts. Gradient enhanced inverse-detected HSQC NMR spectrum;
catalyst hydration levels [53]. High concentrations of H3PO4, ranging Red is MTW, Black is SPA; Regular 1H and 13C NMR spectra are plotted on top of hor-
izontal and vertical axes, respectively. Adapted with permission from Ind. Eng. Chem.
between 110 and 120%, were shown to be optimal, similar to that
Res. 55 (2016) 9140–6. Copyright 2016 American Chemical Society. (For interpretation
observed in the liquid H3PO4 system [35], with the pyrophosphate
of the references to colour in this figure legend, the reader is referred to the web version
phase concentration increasing with H3PO4 concentration [46] and of this article.)
calcination temperature [46,49]. Due to a lack of spatial constraint,
branching is quite significant in distillate products, reflected in cetane relatively high quantity of Type III isoolefinic alkenes [66]. Due to the
numbers of hydrogenated distillate near 30 [54]. Additionally, as number of products formed during acid catalyzed oligomerization
temperature is increased to increase selectivity to distillate, hydrogen processes and the difficulty of separation, significant efforts have oc-
transfer reactions yielding paraffins and aromatics occur. curred to better understand the catalyst-oligomeric product structure-
A new version of SPA, which contains BPO4 in addition to the above property relationship. Characterization of the hydrocarbons formed has
silicon phosphate phases, is being marketed in China as T-99 and ap- utilized advanced GC–MS [67], GCxGC [68], and NMR [66] techniques
pears to be of lower activity than traditional SPA, possibly due to di- to determine product speciation (Fig. 3), where up to 48 compounds in
lution of the active phase [50]. the C8 region have been molecularly identified [69], including cyclic
Isobutene is typically the most reactive olefin over SPA, with n- alkanes [68].
butenes more reactive than propene [41,55]. When pentenes are in- In addition to water tolerance and subsequent corrosion concerns
cluded, methylbutenes are more reactive than n-butenes, and n-pen- from H3PO4, one of the difficulties with SPA, and the typical reason
tenes are the least reactive species [56]. Octenes follow the same re- alternative catalysts have been investigated, is the limited lifetime and
activity pattern, with 2,2,4-trimethylpentenes significantly more disposal of spent catalyst [65]. Typically, spent SPA is considered ha-
reactive than linear octenes, in particular 1-octene, due to the phos- zardous waste, but recent efforts have shown that it can be a phos-
phate ester mechanism [52]. Olefin isomerization is typically quite phorus source for fertilizer [70].
rapid and favors Z-olefins as products [57].
Several recent processes using SPA have been developed. UOP has
utilized the high reactivity of isobutene in the InAlk™ process where the 2.2. Acid resins
isobutene in butene containing feeds is dimerized to 2,2,4-tri-
methylpentenes, separated from the unreacted linear butenes and hy- An acidic material frequently utilized as an alternative to SPA cat-
drogenated to isooctane [58]. A portion of the hydrogenated isooctane alysts are the polymeric resins usually sold for ion-exchange such as the
product is recycled back to the reactor inlet to significantly improve Amberlyst line of products [71]. These typically styrene-divinylbenzene
catalyst life [59]. Honeywell UOP has also been working recently on copolymers have sulfonate groups serving as the anion to the H+ ca-
utilizing SPA, amongst other catalysts, for the oligomerization of flui- tion. Different grades have different cross-linking agents in place of the
dized catalytic cracking (FCC) derived olefins as part of integrated divinylbenzene, amounts of cross-linking, or sulfonate group content.
processes and apparatuses [60,61] in the Catolene™ family to produce Another class of these resins are the perfluorinated polymeric resins
gasoline and/or high propylene yields [62]. Exxon has worked on also containing sulfonate groups often sold under the Nafion product
several catalysts, among them SPA, for the oligomerization of propene line [72]. The chemical structures of both types of polymeric material
to nonene/dodecenes (trimer/tetramer) as the intermediates to produce are illustrated in Fig. 4. Discussion here focuses on recent developments
alcohols and aldehydes [63,64]. SPA produces a particular branch in this active area as Antunes, et al. recently reviewed the use of these
pattern and molecular weight (MW) distribution that is desired in the materials in oligomerization for fuels production [73].
product utilized as carbonylation feed [64,65]. The highly acidic catalysts, with proper control of polar compounds
Though branching, MW, and other properties are dependent on feed such as water [74,75], have been most successfully utilized at low
and process conditions, SPA gives a product with significant methyne temperature (0–100 °C) to selectively oligomerize highly reactive ole-
(CH) and methylene group (CH2) content also corresponding to a fins such as isobutene to the dimeric 2,4,4-trimethylpentene product
which is then hydrogenated to isooctane as in the Neste NExOctane™

84
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Fig. 4. The chemical structures of A) sulfonated styrene – divinyl benzene copolymers


and B) tetrafluoroethylene – sulfonated perfluorovinylether polymers. Both representa-
tions are somewhat simplified, but give a good idea of the complexity and number of
potential acidic polymers.

and Honeywell UOP InAlk™ processes [1,76]. Recently, Taylor and


coworkers have been developing processes to dehydrate isobutanol
obtained from biological fermentation processes to isobutene [77] and
then oligomerize the isobutene to fuels [78], often targeting the jet
range [79]. The resin catalysts used provide good selectivity to primary
oligomerization products, with GC traces showing only a small number
of compounds produced, unlike SPA and most zeolitic catalysts (Section
2.4) which provide a plurality of products from concomitant oligo-
merization, backcracking and isomerization.

2.3. Amorphous silica-alumina (ASA)

Amorphous silica-aluminas have been used as acidic solids for oli-


gomerization for many years [80], in part because the class of materials
has been around for a significant time, but largely because the materials
are stable to higher temperatures than acid resins or SPA, and hence
regenerable by carbon burn [81]. The reactivity of silica-aluminas has
been known to be dependent on the strength of the Brønsted acidity of
Fig. 5. Surface models of hydroxyls on amorphous silica-aluminas calculated by Chizallet,
the site, but these sites also coke relatively rapidly. Hence, it is the Raybaud and coworkers. Reprinted from J. Catal. 325 (2015) 35–47, with permission
moderately strong acid sites that are responsible for the oligomerization from Elsevier.
reaction after the initial time on-stream [82]. Operating in the liquid
phase at temperatures between about 150 °C and 250 °C, long run conjugate base after deprotonation, hence even PBS can transfer pro-
lengths of multiple months prior to regeneration can be accomplished. tons to probe molecules such as lutidine and presumably catalyze oli-
Oxygenates in the feed cause a higher coking rate as carbonyls are gomerization [95,96]. However, it is impossible to simply utilize luti-
readily converted to coke precursors [83]. Additionally, while run dine (or low temperature CO) as a probe molecule to easily characterize
lengths are shorter, ASA based catalysts are capable of converting feed ASA materials as the interaction of the olefin feed or probe molecules
in the presence of significant quantities of poisons such as CO, thio- results from multiple interactions/contributions [96,97]. They also
phene, NH3 and H2S [84,85]. Silica-aluminas are cheap to manufacture confirmed the long observed effect of water on ASA materials, showing
and can have single pass distillate selectivity approaching 70% from that water helps serve as a proton reservoir and obscures acid site
light olefins, but tend to have high hydrogen transfer rates and hence strength.
produce low cetane value (20–30), relatively aromatic distillate Mixed metal oxides, of which silica-alumina is the most studied
[81,83]. Octane values are typically relatively high, with 92–94 RON example, are well known as acidic solids [29]. Both oligomerization
quite commonly reported [81], and the olefins produced by these and cracking occur over boria-alumina, with true oligomerization fa-
processes have extremely high blending octane values [86,87]. vored in the liquid phase at 150 °C [98].
The relative prevalence of tetrahedral and octahedral Al sites is
related to aluminum content in these materials [88], with higher Al
contents approaching γ-Al2O3 at 25% tetrahedral sites. Low Al content 2.4. Zeolitic materials
materials can approach 100% tetrahedral coordination [85,89]. Alu-
minum content is frustratingly inconsistent in the effect on olefin A special class of silica-aluminas are the crystalline, microporous
conversion level, but does tend to influence selectivity patterns, with zeolites [99]. They have been one of the most investigated class of
high silica content materials showing less cracking selectivity [86,90]. materials for the oligomerization of olefins, partially due to the prior
In addition to traditionally precipitated silica-aluminas, mesoporosity commercial application of MFI zeolitic materials in the Mobil Olefins to
has been introduced by using materials such as MTS, SBA-15, and Gasoline and Distillate (MOGD) and PetroSA Conversion of Olefins to
MCM-41 [91,92]. Researchers have also utilized amorphous materials Distillate (COD) processes [7,100], but also because of the control af-
prepared by methods that would traditionally yield zeolite products; forded the scientist in picking the pore size, connectivity and thus the
these proto-zeolitic solids are also useful for oligomerization [93]. shape selectivity exhibited [101,102]. Over 230 individual structure
Recently, Chizallet, Raybaud, and coworkers have studied many types are now known, with each type afforded a 3 letter designation by
aspects of these materials with a computational approach in an effort to the International Zeolite Association (IZA) Structure Commission [103].
unravel the features that contribute to conversion and selectivity in Zeolites are given proper names when synthesized under different
ASAs. They developed a surface model for ASAs highlighting the di- conditions, for example, UZM-9 [104] or ZK-4 [105], but are classified
versity of potential hydroxyls and established the existence of silicic by structure type, which for these two materials is LTA. The structure
and aluminic pseudo-bridging silanols (PBS-Si and PBS-Al, respectively, types are classified by pore size and connectivity, with 8-membered
Fig. 5) [94]. The acidity of ASAs is dominated by the stability of the ring (8MR), 10-membered ring (10MR), and 12-membered ring (12MR)

85
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Fig. 6. Product oligomer characteristics as a function of


geometric pore surface area showing how pore constraint
changes several important product property markers. Data is
from Nicholas, et al. for MTT, MFI, MTW, SPA catalysts [69].
Arrows in the plot indicate the axis on which the labeled data
series are plotted.

delimited pore channels all frequently encountered with typical size to alter the amount of consecutive reactions observed [111]. Isomer
openings of 4.0 Å, 5.5 Å and 7 Å, respectively. distributions are not necessarily equilibrated even when conversion is
In Fig. 6, several characteristics of olefin oligomers produced over complete as de Klerk showed in studies using propene, 1-hexene, and
zeolite catalysts [69] are plotted against the geometric surface area of mixed Fischer-Tropsch C5eC9 cuts as feed [112]. Reported research
the zeolite pore opening. MTT, MFI and MTW geometric surface areas octane number (RON) produced from these feeds is low, typically below
were calculated using the formula for surface area of an ellipse 80, with good cetane values in the 40–50 range. Eni oligomerized
(π*(0.5*A)*(0.5*B)) where A and B are the pore openings given in the cracked naphthas containing mixed C5 and C6 olefins to distillate fuels
IZA database [103]. MTT, MFI and MTW had pore openings of 18, 23, with cetane values around 55 [111]. At Honeywell UOP, we showed
and 26 Å2 respectively, moving from highly constrained to less con- that mixed butene feeds including isobutene achieved higher RON va-
strained pore openings. Solid phosphoric acid was assigned a value of lues of 95, but with poorer cetane of 36 [113].
50 as an unconstrained solid for comparison. High constraints from Most of these studies were performed with relatively high percen-
pore walls (smaller pore openings) lead to more linear products. tages of olefin in the feed, including the COD and MOGD processes, but
Branching in the C8 fraction, whether as average branch number (red de Klerk and coworkers have recently utilized MFI to oligomerize, in
open circles) or the trimethylpentene percentage in the C8 fraction the presence of poisons such as CO, dilute (as low as ∼7%) C2eC4
(blue triangles), increases significantly as pore constraint decreases. At containing Fischer-Tropsch (FT) tail gas feeds to fuels. Interestingly, no
the same time, the linear olefin percentage in the C8 fraction decreases evidence was found for reaction of CO with the olefins, nor for an effect
exponentially (gray open boxes). The cetane number of the distillate on olefin conversion [114].
boiling point range product produced (purple diamonds), a bulk Researchers from Eni have also recently shown that deactivation for
property which is dependent on branching, with linear molecules pre- any particular MFI is dependent only on the amount of olefin processed
ferred for high cetane number, decreases as pore constraint decreases. although different MFI catalysts have different deactivation rates
Effects on branching thus clearly occur throughout the molecular (Fig. 7) [111]. One reason is that crystal size of the MFI zeolite utilized
weight range, not just in the more readily characterized C8 region. Ef- has been repeatedly shown to impact performance, with smaller crys-
fects of pore constraint are also visible via the substitution around the tals yielding longer lifetimes and lower coking tendencies [115,116].
double bond, with the percentage of C8 olefins having 3 substituents Recently, mesoporosity has been shown to significantly impact lifetime,
(out of 4 possible) on the double bond decreasing as pore constraint one of the main drawbacks to utilizing a three-dimensional material
decreases (yellow circles) [69]. versus the one-dimensional materials described later [106,117]. Thus,
what is important is shortening the diffusion path length, whether by
2.4.1. MFI making small crystals or introducing mesoporosity, so that oligomeric
Since the progress shown by Mobil in commercializing the MOGD products do not become coke.
technology, the highest number of studies have been carried out with Acid site density is related to the SiO2/Al2O3 ratio and has been an
catalysts based on the 3-dimensional, 10 membered ring MFI zeolite area of interest for some time [118,119]. Regardless of whether alu-
[106]. In addition to traditional methanol sources, coal has also been minum gradients exist in the MFI zeolite, high aluminum content, and
utilized as the source of syngas, allowing complete coal to gasoline hence high acid site density, leads to significant coking and reduced
processes [107]. PetroSA has also commercialized and is practicing a lifetime [120,121]. The proximity of acid sites is correlated with aro-
MFI based oligomerization process they call Conversion of Olefins to matic molecule formation [121]. By looking at the distribution of di-
Distillate (COD) [100,108]. As discussed above, MFI based catalysts methylhexenes within the C8 fraction, it becomes clear that significant
produce low degrees of branching, primarily methyl in character, fractions of catalysis actually occur on the outside surfaces of the
making it a good catalyst with which to produce distillate range fuels crystals for high acid site density materials. This effect is likely due to
[109]. significant coking of internal channels [120]. An approach taken to
Feed and process conditions both impact fuels properties. increase shape selectivity provided by MFI zeolite has been to deacti-
Temperature significantly affects the amount and type of olefin ad- vate the surface acid sites by titration with an amine. While this ap-
sorbed on the catalyst, hence ethene and propene show decreasing proach hurts overall distillate yields compared to the unmodified MFI,
conversion as temperature increases, while butenes have increased the resulting distillate product can be nearly linear [122].
conversion at higher temperatures [110]. Space velocity can be utilized Care must be taken in these studies as the zeolite crystallite size can

86
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Control of the oligomer product branching pattern has been one of


the driving factors behind use of the MFS (ZSM-57) zeotype, another
10 × 8MR structure, in oligomerization [131]. MFS exhibits a higher
dimer selectivity than other zeotypes during 2-butene oligomerization,
including beta and FER, and a lack of cracking/reoligomerization re-
actions until about 473 K [132]. The selectivity to dimers can be further
improved to above 90% during butene oligomerization when used in a
catalytic distillation reactor, although selectivity falls at high conver-
sion in autoclave testing [133–135].
The MWW zeotype, containing both 10-membered and 12-mem-
bered rings [136], delivers high conversion during olefin oligomeriza-
tion, often in the form of MCM-22, where influence of silanol groups at
channel mouths in interactions with butenes has been observed
[132,137,138]. Exxon developed, and commercially tested, the
EMOGAS process to replace SPA in CatPoly units [139].

2.4.4. 1-dimensional zeotypes


We have recently investigated, amongst other catalysts, the use of
one-dimensional zeolites such as MTW and MTT for the oligomerization
of C4 and C5 olefins to gasoline and distillate [113,140], as these 1-
Fig. 7. Deactivation of MFI catalyst (lower conversion) is only dependent on amount of dimensional materials have longer lifetimes in oligomerization than
olefin fed over the catalyst and is independent of space velocity. Reprinted from Bellussi, many multi-dimensional materials [106,141]. MTW, a large pore zeo-
et.al., Micro. Meso. Mater. 164 (2012) 127–34, with permission from Elsevier. lite containing a relatively small 12MR, shows good selectivity to ga-
soline, particularly once treated with phosphates [142]. Selectivity to
increase significantly as the acid site density decreases unless synthetic distillate range products is 15% prior to phosphorus treatment. As the
precautions are taken [119]. With large crystal size, even low acid site phosphorus content of the 80/20 MTW/Al2O3 catalyst increased, se-
density MFI zeolites can exhibit significant fractions of external cata- lectivity to gasoline improved markedly to about 3 wt%, virtually
lysis due to site/pore blocking [118]. For MFI then, each feed olefin has identical to that of SPA, as shown in Fig. 8. At the same time, acidity
a different optimal combination of diffusion path length and acid site changed significantly as Si-O-Al-O-P moieties form, and activity de-
density for optimal catalysis as Corma and coworkers showed in studies clined, even as the binding matrix was converted from Al2O3 into
using propene and pentene as feeds [117]. crystalline or amorphous AlPO4 depending on the phosphate precursor
[140,142].
2.4.2. Other 3-dimensional zeotypes MTT, a medium pore zeolite containing a relatively small 10MR, can
Beta zeolite, an inter-grown mixture of two 3-dimensional 12- show excellent selectivity to distillate range products from butene
membered ring zeotypes [123], has also been frequently utilized for containing feeds, with over 70% yield of C9+ per pass [143]. Jet range
oligomerization. Recently, Lobo observed, as is typical with acid cata- material, the C9eC20 portion of the distillate boiling point range, can
lysts, that olefin positional isomerization is the fastest reaction occur- also be produced selectively by processes where heavy olefins are re-
ring. They also studied diffusional restrictions, concluding that oli- cycled back to the reactor [144]. MTT is quite shape selective, produ-
gomer transport out of the crystals controlled the rate of cing only 23% selectivity to trimethylpentenes within the C8 fraction
oligomerization, which may also explain the inability to form only di- versus 44% for MFI, 59% for MTW, and 75% for SPA at identical
mers using H-beta under any condition [124]. Chang, Jhung, and conditions. Also, as the pore size decreases, the share of multiply
coworkers utilized this property to trimerize isobutene with selectivities branched olefins produced decreases and the cetane number increases
up to 70%, showing that beta gave higher selectivities to trimer than (Fig. 6) [69].
other zeotypes including FAU, another three-dimensional 12MR [125]. Another frequently utilized 10-membered ring zeotype is TON
Conversion stability could be increased in the FAU catalysts by deal- (ZSM-22), a material structurally related to MTT [145,146]. Control to
umination via steaming, but trimer selectivity decreased concomitantly mono-methyl branching with high yields of dimer and trimer products
[126]. Beta has also been used to catalyze 1-pentene oligomerization,
showing, even at 25% conversion, 40% selectivity to trimer (C15) pro-
ducts [127].

2.4.3. 2-dimensional zeotypes


Celik and coworkers have also studied the oligomerization of pen-
tenes using, amongst other zeotypes, the FER structure, a 10 × 8MR
two-dimensional zeotype, as catalyst [128]. FER showed very different
reactivity patterns, with higher selectivity to decene of 74%, than the 3-
and 1-dimensional zeotypes also tested at 200 °C. This reaction pattern
was ascribed to a lack of bimolecular cracking reactions, behavior also
observed by Huber and coworkers in 1-butene oligomerization [129].
Trimers of isobutene were obtained selectively over FER catalysts at
low temperatures of about 70 °C, with trimer selectivity increasing to
about 70% with increasing isobutene conversion at these temperatures
due to a lack of cracking [130]. GCxGC techniques showed branching in Fig. 8. Effect of phosphorus on oligomerization of C3/C4 feed at 65% conversion showing
the octene dimers produced increases with increasing butene conver- the effect of phosphate treatment on 80/20 MTW/Al2O3 zeolite based catalysts in com-
sion, with product molecular weight following a Schultz-Flory dis- parison to reference solid phosphoric acid (SPA). Reprinted from Nicholas, et.al., Appl.
tribution [129]. Catal. A (2017), with permission from Elsevier.

87
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

and low tetramer formation are hallmarks of ZSM-22 [147]. Exxon also known as catalysts. In a recently reported FeCl3 containing tetra-
researchers have shown that small levels of water in the feed improve alkyl ammonium ionic liquid system, isobutene appears to be oligo-
lifetimes of these catalysts [148,149], while still maintaining the merized to dimers and trimers via Brønsted sites. As for most acids,
branch structure desired in the nonene/dodecene product, used as a cracking reactions were favored at higher temperatures, in this
feed for hydroformylation, although some oxygenates are detrimental case > 60 °C [168].
to lifetime [150]. The introduction of mesoporosity by first desilicating Previously, alkylsulfonate imidazolium triflate ionic liquids had
with NaOH, then dealuminating with oxalic acid appears to increase shown utility for oligomerization of olefins at 90 °C in batch reactions.
lifetime for TON materials and can result in increased distillate se- Isoolefins such as isobutene were more reactive than 1-butene, and
lectivity [151]. Exxon researchers have also sometimes utilized TON in significantly more reactive than long chain mono-olefins. Selectivities
combination with MFS or MFI to deliver the desired branching structure to dimers/trimers of up to ∼90% were achieved at conversions greater
[152,153]. Process conditions including relative amounts of the cata- than 60% [169].
lysts, temperature, and WHSV control selectivity to mono-, di-, and tri- 1-Butyl-3-methylimidazolium tetrachloroaluminate ionic liquids
branched products. Another method utilized to control branching were also shown to oligomerize C2eC8 olefins to dimers and trimers
during oligomerization with TON or MTT is selectivation of external [170]. Yields up to 67% were achieved and decreased monotonically as
acid sites with a bulky amine, ensuring only internal, shape selective the olefin chain length increased. Longer chain lengths were preferred
acid sites are available for reaction [154]. on the imidazolium for activity, presumably for greater lipophilicity
When control of branch structure has not been a significant concern, [169], so the decrease in oligomer yield as the olefin chain length in-
the large, 12-membered ring MOR material has been utilized creases observed here could be due to a mismatch in product/ionic li-
[155,156]. High SiO2/Al2O3 materials showed good selectivity to dimer quid properties.
product and preferential oligomerization of isoalkenes during pentene
and light cracked naphtha conversion [127]. 3. Metal catalyzed oligomerization

2.5. Other non-zeolitic acidic solids Due to the industrial importance of the production of linear alpha
olefins (LAO), a reaction often carried out using homogeneous catalysts,
In addition to the silica and alumina based materials described chemists have “traversed the periodic table searching for new transition
above, other acidic solids including clays, sulfated metal oxides (SMO), metal and lanthanide-based olefin polymerization systems” and a sig-
and polyoxometalates (heteropolyacids) can catalyze oligomerization. nificantly greater number of systems than can be discussed here have
Classically, clays, whether pillared or acid treated, have been used for been explored [171]. The market for LAOs is currently over 5 million
oligomerization, although most show a mixture of oligomerization, metric tons with about 40% of the LAO market used for lubricants,
cracking and cyclization reactions [157–159]. Due to their lower detergents (sulfonated linear alkylbenzenes), oilfield chemicals and
acidity, they are usually regarded as having lower activity than zeolites other fine chemical products, while the other 60% is utilized as co-
and silica-alumina, but can survive higher quantities of poisons in the monomer during polymerization of ethene and propene [172].
feedstream [160]. Metal based catalysts typically utilize a Cossee-Arlman 1,2-insertion
Due to their high acidity, sulfated and tungstated metal oxides show mechanism [173], shown in Fig. 9A, to catalyze olefin oligomerization.
high activity, with good selectivity to true oligomerization products as Following coordination, the olefin inserts into the metal-hydride or
long as the temperature of reaction is kept low enough to avoid metal-alkyl bond with the least substituted carbon bonded to the metal.
cracking. Typically, high selectivity to trimerization of isobutene with This is then typically followed by β-hydride elimination to yield a vi-
some tetramer production is observed over sulfated titania and tung- nylidene substituted product. Significant steric interactions with the
stated zirconia [161,162]. The method of preparation of materials, alkyl chain on the incoming olefin inhibit coordination of the olefin
particularly how the anion modification/sulfation is carried out, has a prior to insertion, so many catalysts only oligomerize ethylene or 1-
large impact on catalytic properties, perhaps due to the varying ratio of olefins. For this reason, metals typically have significantly higher re-
Lewis to Bronsted acid sites, with higher Lewis site density giving activity towards ethene than substituted olefins, so much work has
higher stability [163]. Leading data, for both the SMO and boria-alu- focused on oligomerization of ethene, with some studies also using
mina mixed metal oxide systems, suggest that Lewis acid sites may be propene [174]. Work resulting in polymerization of the olefin feed is
capable of acting as the catalytically active site [98,164]. not discussed here. Mis-insertion of olefin to give 1,2-insertion followed
Polyoxometalates, also referred to as heteropolyacids, have his- by 2,1-insertion as illustrated in Fig. 9B is rare, but possible as is the
torically found utility in olefin oligomerization but had not seen much opposite case of 2,1-insertion followed by 1,2-insertion (Fig. 9C). These
development recently until Iglesia and coworkers studied them to ex- two instances yield linear internal olefins and methylbranched vinyli-
pand their study of deprotonation energy as a descriptor of oligomer- denes respectively. This author has not seen a non-polymerization
ization rate [32,165]. A wide range of elements have been studied as catalyst system performing regioregular 2,1-insertion as illustrated in
the central atom in the Keggin structure with Si and P the two most Fig. 9D to yield methylbranched internal olefins, but this reactivity
frequently studied. With the lowest deprotonation energy, and thus pattern should be possible. Regioregular insertions are normally the
weakest conjugate base, the P substituted Keggin has consistently been products, and while chirality is not reported in any of the products
shown to be the most active heteropolyacid, whether tested as reviewed here, it is worth noting that insertion of substituted olefins
H3PW12O40 or as the Al salt thereof [32,166]. Oligomerization of pro- creates a prochiral metal-alkyl chain. In addition to β-hydride elim-
pene over various salts of phosphotungstic acid gave liquid distribu- ination, chain transfer to monomer, hydrogen, or aluminum alkyls can
tions centered around trimers, with cetane numbers in the range of be observed as growth termination modes [175].
35–40 for Al salts [166]. The preferential oligomerization of isobutene
in the presence of linear butenes over silica-supported tungstosilicic 3.1. Aluminum
acid also yielded a mixture of oligomers (dimers, trimers, tetramers,
etc.) at 60 °C [167]. The largest difficulty with polyoxometalates has Ziegler’s discovery of the Aufbau reaction 65 years ago was the first
consistently been the production of heavy oligomers and coke pre- indication that metals could catalyze the formation of CeC bonds from
cursors. olefins [8]. The insertion of ethene into an aluminum alkyl gives a
pseudo-Poisson distribution of even numbered carbon olefins and is
2.5.1. Ionic liquids used as the primary method of producing linear alpha olefins (LAO)
In addition to the solid catalysts utilized above, liquid systems are [176,177]. Both the Gulfene [178,179] (now Chevron Phillips

88
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Fig. 9. 1,2- and 2,1-insertion mechanisms of an


olefin into a metal-hydride or metal-alkyl bond
leading to longer chain olefins after beta-hydride
elimination to regenerate the hydride. As illustrated
by the red box in part A, almost all systems utilize
regioregular 1,2-insertions. (For interpretation of the
references to colour in this figure legend, the reader
is referred to the web version of this article.)

AlphaPlus®) and Ethyl [180] (now Ineos) processes utilize solution for oligomerization. Davis and coworkers recently synthesized Zn-
based processes with trialkylaluminum as the key propagating species. containing silicates CIT-6 (BEA* zeolite topology) and MCM-41 and
In addition to these well-developed solution phase processes, hetero- showed that exchange of Ni2+ into these materials led to higher se-
geneous catalysts with surface Al-H moieties are also catalytically ac- lectivity to direct propene oligomerization than materials made by
tive for ethene insertion and chain growth as we showed recently exchanging Ni2+ into Al containing materials due to a significant re-
[181,182]. duction of acid catalyzed side reactions [189]. As also observed by Bell
and coworkers in comparisons of Ni2+ exchanged FAU and MCM-41
materials [190], the mesoporous silicates gave more stable conversion
3.2. Nickel profiles [191]. Bell also showed that Ni2+ migrates to the largest space
in the FAU zeolite to yield higher stability. It appears that residual acid
While a number of metals have been utilized for olefin oligomer- sites in these materials catalyze isomerization of product olefins in the
ization, the most frequently utilized catalysts contain Ni, including the zinc materials and cracking/isomerization in Al materials. This beha-
homogenous catalysts utilized in the well-known Shell High Olefins vior can be utilized, as discussed below in the bi-functional oligomer-
Process (SHOP) [183,184]. The general structure of the SHOP catalyst ization section, but is often a detriment here.
is shown in Fig. 10 [185]. Because of the ability to fine tune the Another method of controlling residual acid sites in supported
structure-property relationships using small modifications of the li- nickel catalysts is to over-exchange the sites such that some nickel is
gands in the SHOP type Ni complexes, a number of complexes have wasted as NiO on the support surface [192] as we showed recently in
been investigated and recently reviewed [186,187]. conversion of dilute ethylene to liquids using silica-alumina supports
Coverage here focuses on newer discoveries in this active field and [193,194]. Fluorotetrasilicate micas have also been shown to be useful
is briefer than historically justified given that the utility of nickel in supports for ethylene oligomerization, with intercalation of fluoro-
oligomerization has been known since early investigations of the containing α-diimine ligands into Ni-exchanged micas yielding Schultz-
Aufbau reaction, with both homogeneous and heterogeneous catalysts Flory distributions of even carbon number oligomers with alpha values
recently investigated [188]. The biggest difficulty in preparing het- of between 0.55 and 0.65 for highly active catalysts [195].
erogeneous Ni catalysts for oligomerization has been preparing mate- Dinca and coworkers avoided the use of acidic supports, exchanging
rials with isolated Ni sites, which have been shown to yield active sites Ni2+ onto the Zn2+ cation site in a tris-pyrazolylborate based metal-
organic framework (MOF) framework in order to achieve 95% se-
lectivity to butene with just under 5% selectivity to hexenes at 25 °C,
30 bar ethylene after activation of the catalyst with methylalumoxane
(MAO) [196]. MAO yields Ni-CH3 species, and the catalyst has been
shown to produce butene by the expected Cossee-Arlman 1,2-insertion
pathway [197]. Pyrazolyl based ligands on Ni also show high selectivity
to butene in analogous homogeneous systems [198]. Selectivity to bu-
tene is higher at higher ethylene pressures, suggesting that chain
transfer to monomer rather than β-hydride elimination drives the se-
lectivity pattern for these catalysts.
Ni-diimine catalysts synthesized by post-synthetic modification of
Fe-MIL-101 showed turnover frequencies (TOFs) of ethene up to
∼10.5 × 103 mol (mol Ni h)−1 with selectivity to 1-butene > 90%
[199]. The hydroxyl groups of the Zr6 nodes in the MOF NU-1000 have
also served as a scaffold on which to site-isolate Ni. After activation
Fig. 10. Conceptual pictogram of Ni based SHOP-type catalysts during oligomerization.
with diethylaluminum chloride, the catalyst produced largely butene
Reprinted with permission from Ref. [185].

89
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

systems for dimerization of ethene to 1-butene using Ti(OBu)4 and


trialkyl aluminum [209]. Shortly thereafter, Toyo Soda developed the
Toso dimerization process using phenoxytitanium complexes to achieve
up to ∼85% selectivity to butenes [210]. Titanium alkoxides are the
catalytic precursors still utilized in combination with aluminum alkyls
to generate linear butenes on large scale today, including in the Al-
phaButol® process [205].
Classical thought for these catalysts on the origin of high selectivity
to 1-butene followed a metallacycle mechanism (see Section 4). Recent
work utilizing thorough labeling studies, DFT calculations, and altera-
tions of the aluminum alkyl chain transfer agent have been used to
confirm the reaction mechanism as actually that of Cossee-Arlman (1,2-
insertion) with a high ratio of chain transfer to polymerization
Fig. 11. Chain transfer to monomer in zirconium catalyzed oligomerization reactions.
Reprinted with permission from J. Am. Chem. Soc. 118 (1996) 11670–1. Copyright 1996 [211,212].
American Chemical Society. Bercaw, Labinger and coworkers have recently utilized phenox-
yimine (FI) based ligands forming a tridentate ONO coordination en-
and octene at low ethene conversion at 45 °C, 2 bar [200]. vironment on (FI)Ti(CH3)3 activated with B(C6F5)3 to yield primarily 1-
hexene and C10 olefins at 25 °C, 1 atm C2H4 [213]. Catalyst decom-
position is relatively rapid, but supporting the catalyst on MAO coated
3.3. Group IV: titanium, zirconium and hafnium silica significantly increases lifetime while retaining selectivity to 1-
hexene and branched C10 olefins [214]. C10 olefins are generated from
The next most commonly utilized metal has been zirconium, well- cotrimerization of a 1-hexene with 2 ethenes, so this ligand set may
known in olefin polymerization catalysis [201]. To produce oligomers actually oligomerize olefins through a metallacycle mechanism rather
rather than polymers, chain transfer agents such as trialkylaluminums than the typical 1,2-insertion mechanism. Related ligands with an N-
or inhibitors such as thiophene or arylethers are utilized to control the heterocyclic carbene in place of the imine such that an OCO co-
molecular weight of the product. Multiple commercialized LAO pro- ordination environment is achieved have been used for Zr and Hf
cesses including those from Idemitsu [202,203], the Alpha-Sablin pro- complexes for trimerization of 1-hexene [215,216]. Upon activation of
cess from Sabic/Linde [204], and Alphaselect process from IFP-Axens the benzyl derivative (OCO)MBz2 with HN(CH3)2Ph+B(C6F5)4−, the Zr
[205] utilize this general concept. They run at moderate pressures of complex is 77% selective to C18 compounds at room temperature in
20–40 bar and relatively low temperatures of about 25–100 °C. Zr based liquid hexene. As observed previously, the Hf derivative is considerably
catalysts typically provide Schultz-Flory distributions of product, slower [215].
though via the use of high concentrations of chain transfer agents
pseudo-Poisson distributions are possible. Work on the mechanism of
reaction shows that instead of β-hydride elimination, chain-transfer to 3.4. Group V: vanadium and tantalum
monomer can drive these narrow carbon number distributions of olefin
oligomers (Fig. 11) [206]. Moving over a group, homogeneous vanadium complexes chelated
Hessen and coworkers also showed that the insertion of ethene into with nitrogen based ligands [217] have shown high (> 90%) selectivity
the Hf complex is 40 times slower than the same Zr complex, leading to to 1-butene or 1-hexene during ethene oligomerization [218,219]. Both
low overall productivity [206]. Slow insertion of propene into Hf-CH3 vanadium (V) and (III) complexes (Fig. 12) have been reported for
versus Zr-CH3 was observed by Teuben and coworkers in a [Cp*2M oligomerization of ethylene to Schultz-Flory distributions of olefins.
(CH3)2(tetrahydrothiophene)]+ [BPh4]− catalyst system [207]. The Zr The size of imido substituents in vanadium (V) complexes plays a sig-
catalyst produced a range of oligomers from C6 to C24, but Hf selectively nificant role in controlling the oligomerization, with smaller sub-
produced only one isomer of C6 and C9. stituents giving rise to polyethylene instead of oligomers [220].
In addition to the homogeneous systems described above, hetero- In one of the few studies where both ethene and a larger olefin were
genous zirconium catalysts have been utilized for oligomerization re- tested, V(III) complexes were shown to have significantly lower re-
actions. Zr-H systems prepared by reacting organometallic precursors activity with propylene than with ethylene [218]. Highest productivity
with hydroxyls of a silica-alumina support followed by H2 treatment to was observed with X = Cl while still achieving high 1-hexene purity
hydrogenolyze remaining Zr-C bonds dimerize 2-butene quite effec- [219]. The steric bulk of the ligands also affects 1,2- versus 2,1-inser-
tively to mono- or di-branched octenes [208]. tion of the second propylene showing again that steric interactions,
Prior to the development of the Zr/Hf based catalysts, titanium was both of the ligands and of the incoming olefin, play a significant role in
a key metal for oligomerization. Ziegler reported one of the first catalyst selectivity to oligomers in these group IV and group V catalysts. All the

Fig. 12. Vanadium complexes utilized in ethylene


oligomerization. A) vanadium (V) complexes, B)
vanadium (III) complexes [217–220].

90
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

phenanthroline ligands showing some interesting selectivity patterns


towards oligomers during ethene oligomerization [232]. Recently,
Cariou, et.al., at BP showed iron-bipyridyl complexes in combination
with ZnEt2 and the aforementioned Fe-PDI type catalyst are capable of
regenerating the ZnEt2 chain transfer agent and provide for a single-pot
ethene oligomerization system. Higher quantities of ZnEt2 lead to
higher selectivity to oligomerization, but at a cost of product α-olefin
purity [233].

4. Metallacycles for ethylene dimerization, trimerization and


tetramerization

The selective conversion of ethene to butene, hexene and octenes


has been extensively investigated since the advent of the nickel cata-
Fig. 13. Control of ethylene oligomerization in homogeneous catalysts by metallacycle
lyzed SHOP technology [234,235]. These processes are often catalyzed
formation. Reprinted with permission from Chem. Rev. 111 (2011) 2321–41. Copyright
2011 American Chemical Society.
by homogeneous, well-defined molecular catalysts with intricate
structure property relationships, some of which were discussed pre-
vanadium complexes are first activated with MAO, a reaction that has a viously in the metal catalyzed oligomerization Section 3.
complicated, time-resolved impact on oligomerization activity and It is often difficult to determine whether a particular catalyst system
performance [221]. V-alkyl and/or -hydride species appear to be utilizes a metallacycle mechanism as drawn in Fig. 13, a combination of
formed during this activation step which are responsible for the cata- metallacycle and 1,2-insertion (Cossee-Arlman) mechanisms (i.e. the
lytic activity [222]. Supporting these complexes on supports results in relative rates are competitive) or simply the 1,2-insertion mechanism as
significantly increasing product molecular weight as we have observed discussed above in Section 3 for metal catalyzed oligomerization. Im-
previously [223,224]. plicit in the metallacycle mechanism is a two electron oxidation and
Tantalum has also been utilized, with Ta-H systems prepared as reduction of the metal center upon formation and elimination, respec-
above for Zr-H showing a mixture of dimerization and metathesis be- tively, of the metallacycle. Using detailed labeling experiments, often
havior at 150 °C [208]. Metathesis is favored at higher temperature, coupled with theory, it is possible to determine that most catalyst sys-
with oligomerization favored at lower temperatures as we recently tems based on zirconium, titanium, iron, nickel, and vanadium utilize
showed [225]. Grafting Ta(CH3)3Cl2 onto partially dehydroxylated si- the 1,2-insertion mechanism, but eliminate primarily to a single pro-
lica gave a species that, upon exposure to ethene, reduces to a TaIII duct for one reason or another [212,236,237]. Chromium based cata-
intermediate with two chloride ligands and one siloxy bridge to the lysts appear to be a significant exception as noted by McGuinness [238].
surface [226]. This species is trigonal bipyramidal and selective to 1- A large number of ligands have been utilized with chromium based
hexene during ethene oligomerization. Thermal treatment of the silica catalysts to control activity and selectivity to 1-olefins, in particular 1-
supported Ta-H catalyst above leads to a silica-supported TaIII species hexene [239]. Building off of discoveries made at Union Carbide [240],
which polymerizes ethene, a behavior Basset and coworkers showed CP Chem commercialized the Phillips Trimerization System which
was due to a square pyramidal coordination environment [227]. utilizes a Cr based catalyst with a pyrrolide ligand(s) and aluminum
alkyl activation/chain transfer agents [241,242]. While significant de-
bate still occurs as to whether the catalyst proceeds via a Cr(II)/Cr(IV)
3.5. Cobalt- and iron-based catalysts or Cr(I)/Cr(III) catalytic cycle [243,244], 2nd order dependence of
ethene pressure and theoretical work show that a metallacycle me-
In an attempt to produce linear oligomers from 1-olefins, Hermans, chanism is at work [245]. Branched C10 olefins appear to be an inherent
Huber and coworkers utilized cobalt oxides supported on carbon sup- side product due to co-trimerization of 1-hexene with ethene [246]. In
ports to dimerize 1-butene to linear octenes, achieving up to ∼80% addition to studies of the mechanism, significant work on synthesis and
selectivity to linear octenes [228]. Selectivity to linear products is de- repeatability of the catalyst preparation have also occurred [247]. Ac-
pendent on conversion, with higher conversion leading to more bran- tivities up to about 106 g hexene mol−1 Cr h−1 have been achieved. A
ched products. Linear octenes could be formed from 1-butene by 1,2- complex generated using tetramethylpyrrole showed 95% selectivity to
insertion followed by 2,1-insertion as shown previously in Fig. 9B, or by 1-hexene with an activity of 7 × 105 g mol−1 Cr h−1 in methylcyclo-
collapse of a 1,4-disubstituted metallacycle (Section 4). Hermans, hexane solvent [248]. An important point from all these studies is that
Huber and coworkers proposed cobaltocyclopentanes as the catalytic hemi-labile halide and/or ether ligands are an important component to
intermediate and noted the mixed oxidation state oxide Co3O4 was the achieve selectivity to 1-hexene.
more active precatalyst [228]. Previously, in a series of papers, Schultz, Tridentate ligands containing a nitrogen and two additional soft
et.al., had utilized similar catalysts to dimerize propene, butene, hexene donors have also been well studied. PNP ligands coordinating through 2
and ethene, invoking a Co-H species as the key intermediate [229–231]. sterically unhindered phosphines and an amine were some of the first
While the reaction mechanism and catalytically active species remains utilized [249]. Recent theoretical work shows that substitution on the
under debate, it is clear that these cobalt oxide on carbon catalysts o-carbon of the arylphosphine increases selectivity to hexene by en-
provide higher selectivity to linear products than most other oligo- suring single ethene insertion into the chromacyclopentane [250]. In
merization catalysts. this system as well, hemilabile methoxy and chloride ligands are quite
While better known as highly active olefin polymerization catalysts, important to the trimerization selectivity [251,252]. Researchers at
a number of iron complexes, often in combination with chain transfer Sasol have recently calculated a cone angle based parameter for the
agents, have shown utility for oligomerization [171]. A number of bi- steric bulk around the N center of a family of PNP ligands and shown
siminopyridine (PDI) ligands, forming either Fe or Co complexes, have that intermediate steric bulk at the N leads to higher selectivity to
excellent dimerization selectivities for 1-butene feed at low tempera- tetramerization to 1-octene [253]. Thus, a complex synthesized with a
ture. The mechanism reported here is Cossee-Arlman, with a 1,2-in- ligand having an isopropyl substituted N gave 70% selectivity to 1-
sertion into the metal hydride then often followed by a 2,1-insertion to octene, while n-propyl, t-butyl, or cyclohexyl (Fig. 14) only allowed
obtain a linear dimer followed by β-H elimination [171]. A number of selectivity of about 60% [253,254]. At 45 °C, the isopropyl substituted
variations of this ligand have been synthesized, with 1,10- ligand was also extremely active, converting up to 5.9 × 106 g ethene

91
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Fig. 15. 1-hexene trimerization intermediates. The three chromacyclopentanes possible


into which the third insertion can take place. Reprinted with permission from ACS Catal.
6 (2016) 3008-16. Copyright 2016 American Chemical Society.

coupled with production of polymer, both of which were concomitant


with CrII formation by reduction of CrIII.
Fig. 14. Molecular structure of [Cr(PNP)Cl2(μ-Cl)]2. Reprinted with permission from J.
Am. Chem. Soc. 126 (2004) 14712-3. Copyright 2004 American Chemical Society.
5. Multifunctional metal and acid catalyzed
−1 −1
g Cr h .
While many metal catalysts, whether utilizing metallacycle or
SNS ligands coordinating through 2 thioethers and an amine also
Cossee-Arlman mechanisms, have been used, catalysts containing both
showed upwards of 97% selectivity to hexene (99.6+% 1-hexene in the
metal and acid sites have been used extensively, particularly for the
C6 fraction) with TOFs at 85–90 °C approaching 300,000 [255]. Pyr-
oligomerization of ethylene, which is difficult to oligomerize with only
rolidine containing NNN ligands exhibit selectivity dependence on the
acidic sites [265]. Most multi-functional catalysts studied use Ni as the
amount of MAO utilized in activation at 80°, 20 bar ethene, with
metal site with a variety of solid acids [188]. IFP’s development of the
500–1000 equivalents Al/Cr leading to > 90% selectivity to mixtures of
Dimersol and Difasol processes provided one of the first multi-func-
C4eC8 olefins [256]. Dearomatized PNP variants on pyridine based
tional homogeneous systems comprising both dimerization and iso-
pincer complexes showed moderate activity and also gave > 90% se-
merization [1,205]. The cationic nickel complexes utilized as the
lectivity to mixtures of C4eC8 olefins. The aromatic PNP or PNC (N-
AlCl4− salt catalyze dimerization of propene or butene feeds to iso-
heterocyclic carbene) variants instead yielded polymer [257].
hexenes or octenes, respectively [266,267]. The octenes are feed for
A family of ligands containing N-heterocyclic carbenes with P do-
isonanol production [268]. Difasol incorporated tetrachloroaluminate
nors have also shown utility in oligomerization of ethene. Square-pyr-
ionic liquids into the process to ease the separation and recycle of the
amidal CrIII complexes with 3 benzyl ligands produced a full range of
nickel catalyst [269,270]. In total, Axens has licensed 35 of these units
oligomers with activities up to 8.8 × 104 g C2H4 g Cr−1 h−1 at 30 °C,
worldwide [271,272]. On the heterogeneous side, Nicolaides and
10 bar ethene after activation with MAO. Use of EtAlCl2 as the acti-
coworkers carried out the classic work using nickel loaded onto
vating agent turned the complexes into polymerization catalysts while
amorphous silica-aluminas to produce fuels [273,274].
square planar CrII complexes were poorly active [258]. Similarly, mixed
UOP has recently also utilized Ni/ASAs in efforts to oligomerize
valence complexes using phosphino-amino ligands show a dependence
dilute (∼25%) ethylene feeds to mixtures of distillate and gasoline
of activating agent on selectivity, with triethylaluminum added to tri-
[275,276]. Even in the presence of significant quantities of impurities
methylaluminum depleted MAO showing excellent selectivity to 1-
including H2, CO, NH3 and H2S, these Ni/ASA based catalysts continue
hexene while other activating agent combinations yield mixtures of
to convert feed [277]. The multi-functional catalysts have significantly
olefins/polymer [259]. Diphosphine (PP) based ligands have also
higher conversion and selectivity to distillate versus the single function
shown utility. A series of silicon bridged phosphines gave up to
ASA base catalysts [84]. Hulea and coworkers have utilized Ni on
4 × 106 g C2H4 (molCr h)−1 with 78% selectivity to 1-octene at 45 °C,
mesoporous silica-aluminas to oligomerize ethene to a full range of
4 MPa C2H4. Activity and selectivity were always to a mixture of C6 and
oligomers with productivities up to 130 g oligomer/g catalyst [278].
C8 with varying percentages depending on temperature, solvent and
Activation temperature/condition of the materials is important as it
pressure [260].
appears Ni+ or dehydrated Ni2+ are the nickel-based active sites [279].
Theopold and coworkers have shown that 2,4-dimethyl-diketimi-
Catalysts comprising Ni on zeolites including MFI have been studied
nate (nacnac) based ligands on chromium produce useful ethene tri-
for a number of years [188,280]. Recently, Ni on MWW materials in-
merization catalysts. CrI dimers with μ-η2:η2 bound N2 gives essentially
cluding MCM-22 and MCM-36 have been used. The more mesoporous
100% selectivity to 1-hexene at 30 °C, 1 atm H2 without need for acti-
Ni-MCM-36 materials exhibit 2–5 times higher activity than the mi-
vation [261]. Triazacyclohexane ligands have been explored by Köhn
croporous Ni-MCM-22 materials [281]. Exchanging additional Ni onto
and shown selective (> 80%, often greater than 90%) trimerization of
the MCM-36 up to 0.28 Ni2+/H+ gave higher average activity in ethene
propene, 1-pentene, -hexene, -octene, or styrene [262,263]. Due to the
oligomerization [282]. The same increase in activity with increasing
substituents on the alkene, mixtures of trimeric compounds are formed,
nickel loading was seen in the Ni-beta system up to the point at which
the ratios of which can be explained by looking at the steric hindrance
NiO was visible by XRD. The Ni-beta catalyst gives non-Schultz-Flory
of insertion of the third olefin into the chromacyclopentane (Fig. 15).
distributions of oligomers with significant branching as one would ex-
1,2-insertions into the 1,3-disubstituted chromacyclopentane appear to
pect for a bifunctional metal-acid catalyst [283].
be favored [262].
The highly acidic sulfated and tungstated metal oxides have also
In addition to the homogeneous catalysts shown above, solids have
been combined with nickel to form bifunctional oligomerization cata-
also been utilized for selective oligomerization of ethene. Cr-MIL-100
lysts. Zong and coworkers supported NiSO4 on γ-Al2O3 containing
was synthesized and evacuated at different temperatures in order to
Fe3O4 to form a Ni-sulfated alumina catalyst for oligomerization of FCC
remove free and bound water molecules or F−/OH− anions co-
cracked gasoline (C5/C6 olefins). Diesel yield was dependent on Ni
ordinated to the Cr centers. After evacuation at 150 °C and activation
loading, with 7 wt% giving the optimum 45% yield [284]. Nickel on
with Et2AlCl, Cr-MIL-100 showed 69% selectivity to C8, 26% to C6 and
tungstated alumina catalyzed isobutene dimerization, showing up to
4% to C10 olefins at 5 × 105 g C2H4/(mol Cr h)−1 activity [264].
80% selectivity to octenes with sol-gel produced catalysts. Reaction at
Higher evacuation temperatures gave higher oligomerization activities
150 °C was optimal for dimer selectivity while the method of catalyst

92
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

preparation significantly affected catalyst stability [285]. Impregnated


catalysts showed decreased selectivity to dimer with increasing tem-
perature (50–150 °C) as had been previously observed for sulfated ti-
tania supported catalysts [286].

6. Oligomerization using combination processes

In addition to catalysts containing two functions, combination


processes where either two different oligomerization catalysts or two
reactions, at least one of which is oligomerization, have also been de- Fig. 16. Use of multiple oligomerization techniques to synthesize fuels. Harvey, et.al.,
veloped to make use of the reactivity patterns of oligomerization. utilize first a metallacycle mechanism, then a 1,2-insertion mechanism to synthesize
distillates. Reproduced from Ref. [304] with permission of The Royal Society of Chem-
istry.
6.1. Oligomerization/Alkylation

While the initial use of solid phosphoric acid was for oligomeriza- in the C8 fraction at a WHSV of 0.38, followed by oligomerization of the
tion, a quickly developed second use was for alkylation of benzene with obtained liquids over Amberlyst-35 to achieve overall yields of ∼42%
light alkenes [287]. Recently, these two uses have been combined in jet fuel [307].
processes where oligomerization of olefins and removal of benzene
from gasoline by alkylation are both goals. In the process, benzene is 6.3. Alternative CeC bond formation/Oligomerization
converted to mixtures of ethylbenzene, cumene, and multiply alkylated
benzenes in a mixture with oligomers [288,289]. Sasol utilized sig- Willauer and coworkers took a different tack, hydrogenating CO2 to
nificant excesses of olefin to achieve up to 80% benzene reduction with C2H4 containing feeds over modified Fischer-Tropsch catalysts and then
SPA catalyst. About 70/30 gasoline/distillate ratios were observed with studying the oligomerization of C2H4 over Ni/ASA bifunctional cata-
propene converted at higher rates than butenes [289]. lysts, achieving 56% yields of jet fuel range materials at 0.15WHSV
At UOP, we have utilized UZM-8 zeolitic materials as the catalyst for [308]. The use of methanol to olefins to produce light olefins for oli-
this process, showing equivalent reaction rates of the olefin with ben- gomerization processes has been commercially practiced, as noted
zene and toluene and slower reaction rates with higher alkylated arenes earlier, but deserves mention here [7,106]. Finally, dimethylether was
including xylenes and mesitylene [290,291]. Using dilute (∼25vol%) converted to jet fuels in a hypothetical process by NREL researchers.
ethylene as the olefin feed, over 20% reduction of benzene was possible Triptene (2,3,3-trimethyl-1-butene) and other C6eC8 branched olefins
in a full reformate (BTX, A9) feed even in the presence of H2, H2S, CO whose carbon skeletons are available from acid catalyzed homologation
and CO2 [292]. of DME [309] were oligomerized to C10eC20 distillate range oligomers
Following on the success of MWW based alkylation processes [293], with up to 52% yields at 80 °C over Amberlyst-35 with some concurrent
multiple researchers have studied MWW materials including MCM-22 isomerization and cracking [310].
for combination oligomerization/alkylation processes for benzene re-
duction. Laredo and coworkers have shown that MCM-22 is sig- 7. Conclusion and future outlook
nificantly more stable and active than beta zeolite catalysts for con-
version of benzene in refinery derived aromatic feeds [294]. As seen After 80 years of work following Ipatieff’s initial discoveries, the
previously, high olefin to benzene ratios are necessary for efficient re- oligomerization of light olefins is still an active research area of com-
duction of benzene and they also observed that higher benzene contents mercial importance. Acidic catalysts such as solid phosphoric acid and
in the aromatic feedstock are beneficial to C6H6 reduction [295,296]. zeolites are utilized in multiple commercial processes and continue to
EMRE, Mobil and Badger workers converted C2eC4 olefins, preferring be under investigation as new material types and processes are dis-
C3eC4 mixtures, over MCM-22 or MCM-49 in the presence of various covered/developed. Significant work has been performed using the
benzene containing refinery streams to provide an alternative benzene most common zeotypes and from that, general rules about stability and
reduction strategy to meet MSATII regulations [297–299]. The product shape selectivity have been developed on the basis of pore size
BENZOUT™ process was scaled up and tested in a refinery [300–302]. and connectivity.
Distillate range fuels, whether jet fuel or for transportation, have Work is still needed as new zeotypes are discovered, partly to ensure
been the target in a number of combination processes. Sasol researchers that the rules continue to hold even for new zeotypes, but also to ex-
used SPA to catalyze oligomerization and alkylation of propene and plore zeolites containing multiple size channels in a single material.
benzene or toluene to make jet fuels containing some aromatic mole- Will oligomerization in a material containing both a 10MR and a 12MR
cules. Oligomerization rates are faster at higher propene/arene ratios channel produce product like that from a 10MR or that from a 12MR?
and alkylation rate faster at low propene/arene ratio [303]. What about oligomerization in a material containing multiple size rings
in a single channel? How linear will the product be from a uni-di-
6.2. Oligomerization/Oligomerization mensional 8MR zeolite? These are the types of questions that remain to
be answered, frequently addressing shape selectivity, but which will
Harvey and coworkers have recently utilized combinations of oli- often require material discovery before the question can be addressed.
gomerization processes to synthesize distillate range materials An area that has held back the development of novel acid catalysts is
[304–306]. Fig. 16 shows an envisioned process where ethylene is the ability to characterize the product produced by the new catalyst.
converted to hexene via one of the metallacycle derived trimerization While GC techniques, including GCxGC, have developed significantly
catalysts. They then utilize Cp*2ZrCl2/MAO to convert 1-hexene to C12 since the advent of the field, additional molecular based characteriza-
and C18 branched molecules via 1,2-insertion pathways [304]. Pre- tion of the products produced is necessary. Today, it is possible to
viously, they had showed 1-butene conversion to 2-ethylhexene via 1,2- characterize, in significant detail, the molecules produced by oligo-
insertion over Group IV catalysts followed by acid catalysis to convert merization in a single carbon number range such as the C8 region.
the C8 molecules to jet fuels [305,306]. However, this same detail is often not available for the C9 or C10 regions
Hwang and Chae have also shown the conversion of ethene to jet of the same product and therefore assumptions are made that dis-
fuels with a combination of processes. They used Ni supported on Al- tribution of the C9 and C10 fraction is similar to that of the C8. For a
containing SBA-15 at 200 °C/10 bar to transform ethene to liquids rich given catalyst, this may be true, or may start to break down as Exxon

93
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

researchers have shown for MFI based catalysts. Additional efforts, both [26] https://www.epa.gov/assessing-and-managing-chemicals-under-tsca/fact-sheet-
nonylphenols-and-nonylphenol-ethoxylates (Accessed 26 May 2017).
to ease the time spent to characterize a particular region of the product, [27] Nexant Inc. ChemSystems Process Evaluation Research Planning (PERP)
and to be able to establish the identities of molecules within that region Report—Non-Phthalate Plasticizers 2014S5, December 2014.
are needed. [28] Nexant Inc. ChemSystems Process Evaluation Research Planning (PERP)
Report—Ethoxylates 2015S1, September 2015.
Metal based catalysts have become of significant importance over [29] G. Busca, Chem. Rev. 107 (2007) 5366–5410.
the last 10 years, particularly with the development of homogeneous [30] A. de Klerk, Ind. Eng. Chem. Res. 44 (2005) 3887–3893.
catalysts capable of selective dimerization or trimerization which [31] H. Pines, The Chemistry of Catalytic Hydrocarbon Conversions, Academic Press,
1981.
achieve high selectivities to 1-butene, 1-hexene, or 1-octene. The next [32] M.L. Sarazen, E. Doskocil, E. Iglesia, J. Catal. 344 (2016) 553–569.
frontier of development here will be to develop selective oligomeriza- [33] V.N. Ipatieff, Ind. Eng. Chem. 27 (1935) 1067–1069.
tion to longer chain oligomers such as C12, C14, or C16, currently [34] V.N. Ipatieff, H. Pines, Ind. Eng. Chem. 28 (1936) 684–686.
[35] A. de Klerk, D.O. Leckel, N.M. Prinsloo, Ind. Eng. Chem. Res. 45 (2006)
available only by fractionating the total product from a full range LAO
6127–6136.
process. Another area where needed development will occur is in the [36] A. De Klerk, J.J. Spivey, K.M. Dooley (Eds.), Catalysis, vol. 23, Royal Society of
production of linear oligomers from substituted olefins, such as the Chemistry, Cambridge, UK, 2011, pp. 1–49.
reasonable selectivity to linear octenes from 1-butene reported in [37] P.A. Maschwitz, L.M. Henderson, Adv. Chem. 5 (1951) 83–96 (Chapter 9).
[38] V. Haensel, V.N. Ipatieff, Ind. Eng. Chem. 38 (1946) 1045–1047.
Section 3.5. Head to tail coupling has been a long sought goal, so [39] D. York, J.C. Sheckler, D.G. Tajbl, UOP Catalytic Condensation Process for
continued development of understanding of the reaction mechanism Transportation Fuels, Chapter 1.3, in: R. Meyers (Ed.), Handbook of Petroleum
and structure-property relationship will help. Additionally, the advent Refining Processes, McGraw-Hill, 1986.
[40] http://www.uop.com/processing-solutions/petrochemicals/detergents/higher-
of heterogeneous catalysts capable of the same tight structure-property olefins/ (Accessed 24 July 2016).
relationships developed in the homogeneous cases would make devel- [41] G.E. Langlois, J.E. Walkey, Petroleum Refiner 31 (1952) 79–83.
opment of selective oligomerization processes even more commercia- [42] M.E. Dry, Appl. Catal. A 276 (2004) 1.
[43] T. Hamamatsu, N. Kimura, T. Takashima, T. Morikita, Solid phosphoric acid cat-
lizable. alyst and method for dimerization of olefin using same, to Nippon Oil Corporation,
In addition to oligomerization only processes, researchers have de- US Patent 7,741,527, June 22, 2010.
veloped multiple processes where oligomerization is one of the steps or [44] N. Kimura, T. Hamamatsu, Olefin dimers and method for producing and washing
olefin dimers, to JX Nippon Oil & Energy Corporation, US Patent 9,314,784, April
key reactions in a process. Impediments here are largely due to the
19, 2016.
necessity to match oligomerization conditions with the conditions uti- [45] T. Hamamatsu, N. Kimura, T. Takashima, T. Morikita, Solid phosphoric acid cat-
lized for another reaction, so advancement of oligomerization catalysts alyst and method for dimerization of olefin using same, to Nippon Oil Corporation,
US Patent 8,203,025, June 19, 2012.
capable of operating under different temperature or pressure regimes,
[46] J.H. Coetzee, T.N. Mashapa, N.M. Prinsloo, J.D. Rademan, Appl. Catal. A 308
or the development of suitably scalable dual batch processes will cer- (2006) 204–209.
tainly help progress in this area. [47] N.M. Prinsloo, Ind. Eng. Chem. Res. 46 (2007) 7838–7843.
Overall, oligomerization has been a commercially useful reaction to [48] T.R. Krawietz, P. Lin, K.E. Lotterhos, P.D. Torres, D.H. Barich, A. Clearfield,
J.F. Haw, J. Amer. Chem. Soc. 120 (1998) 8502–8511.
upgrade low valued light olefins to higher valued products for over 80 [49] T.-H. Chao, F.P. Wilcher, M.R. Ford, A.Z. Ringwelski, Solid Phosphoric Acid
years and will continue to be with the further development of the new Catalyst, to UOP LLC, US Patent 4,946,815, August 7, 1990.
and soon to be discovered advancements detailed within. [50] J. Zhang, Y. Yan, Q. Chu, J. Feng, Fuel Process. Technol. 135 (2015) 2–5.
[51] L. Xu, W. Turbeville, G.A. Korynta, J.L. Braden, Solid phosphoric acid with con-
trolled porosity, to Sud-Chemie, Inc., US Patent 7,557,060, July 7, 2009.
Acknowledgment [52] A. de Klerk, Ind. Eng. Chem. Res. 45 (2006) 578–584.
[53] N.M. Prinsloo, Fuel Process. Technol. 87 (2006) 437–442.
[54] A. de Klerk, Energy Fuels 20 (2006) 439–445.
CPN acknowledges Honeywell UOP for permission to publish. [55] V.N. Ipatieff, R.E. Schaad, Ind. Eng. Chem. 37 (1945) 362–364.
[56] A. de Klerk, D.J. Engelbrecht, H. Boikanyo, Ind. Eng. Chem. Res. 43 (2004)
References 7449–7455.
[57] A. de Klerk, Ind. Eng. Chem. Res. 43 (2004) 6325–6330.
[58] R.L. Mehlberg, P.R. Pujado, D.J. Ward, et al., Handbook of Petroleum Processing,
[1] P.R. Pujado, D.J. Ward, Catalytic Olefin Condensation Chapter 9.2, Handbook of in: S.A. Treese (Ed.), 2015, pp. 457–478.
Petroleum Processing, in: D.S.J. Jones, P.R. Pujado (Eds.), Springer, 2006. [59] L.O. Stine, B.S. Muldoon, S.C. Gimre, R.R. Frame, Process for oligomer production
[2] G. Egloff, Ind. Eng. Chem. 28 (1936) 1461–1467. and saturation, to UOP LLC, US Patent 6,284,938, September 4, 2001.
[3] G. Egloff, E. Wilson, Ind. Eng. Chem. 27 (1935) 917–933. [60] C.P. Nicholas, C.D. Freet, S.L. Krupa, K.M. Vanden Bussche, T.M. Kruse, Process for
[4] G. Egloff, C.I. Parrish, Chem. Rev. 19 (1936) 145–161. Oligomerizing Light Olefins Including Pentenes, to UOP LLC, US Patent 9,567,267,
[5] V.N. Ipatieff, B.B. Corson, G. Egloff, Ind. Eng. Chem. 27 (1935) 1077–1081. February 14, 2017.
[6] V.N. Ipatieff, R.E. Schaad, Ind. Eng. Chem. 37 (1945) 362–364. [61] C.P. Nicholas, C.D. Freet, S.L. Krupa, K.M. Vanden Bussche, T.M. Kruse, Apparatus
[7] C.D. Chang, Catal. Today 13 (1992) 103–111. for Oligomerizing Light Olefins, to UOP LLC, US Patent 9,522,373, December 20,
[8] K. Ziegler, Angew. Chem. 64 (1952) 323–329. 2016.
[9] A. De Klerk, Energy Fuels 19 (2005) 1462–1467. [62] Z. Fei, H. Lim, UOP Catolene™ Process, Chapter 1.4, Handbook of Petroleum
[10] M. Cowley, Org. Process Res. Dev. 11 (2007) 286–288. Refining Processes, in: R.A. Meyers (Ed.), 4th ed., McGraw Hill, 2016.
[11] C.T. O’Connor, M. Kojima, Catal. Today 6 (1990) 329–349. [63] S.W. Beadle, J.S. Godsmark, R.L. Wolf, Oligomerisation of olefins, to ExxonMobil
[12] D.M. Alonso, J.Q. Bond, J.A. Dumesic, Green Chemistry. 12 (2010) 1493–1513. Chemical, US Patent 8,598,396, December 3, 2013.
[13] K. Toch, J.W. Thybaut, G.B. Marin, Appl. Catal. A 489 (2015) 292–304. [64] N.A. de Munck, P.J. Ayre, A.F. da Costa, E. Kooke, Branched olefin compositions,
[14] J.R. Shahrouzi, D. Gillaume, P. Rouchon, P. DaCosta, Ind. Eng. Chem. Res. 47 to ExxonMobil Chemical, US Patent 7,919,551, April 5, 2011.
(2008) 4308–4316. [65] S.H. Brown, J.S. Godsmark, G.M.K. Mathys, Olefin oligomerization process, to
[15] X. Huang, D. Aihemaitijiang, W.-D. Xiao, Chem. Eng. J. 280 (2015) 222–232. ExxonMobil Chemical, US Patent 7,786,337, August 31, 2010.
[16] L. Ying, J. Zhu, Y. Cheng, L. Wang, X. Li, J. Ind. Eng. Chem. 33 (2016) 80–90. [66] C.P. Nicholas, L. Laipert, S. Prabhakar, Ind. Eng. Chem. Res. 55 (2016)
[17] S.M. Stigler, Stat. Prob. Lett. 1 (1982) 33–35. 9140–9146.
[18] G.J.P. Britovsek, S.A. Cohen, V.C. Gibson, P.J. Maddox, M. van Meurs, Angew. [67] R. Bekker, N.M. Prinsloo, Ind. Eng. Chem. Res. 48 (2009) 10156–10162.
Chem. Int. Ed. 41 (2002) 489–491. [68] R. van der Westhuizen, H. Potgieter, N. Prinsloo, A. de Villiers, P. Sandra, J.
[19] G.V. Schulz, Z. Phys. Chem. B 30 (1935) 379. Chrom. A 1218 (2011) 3173–3179.
[20] P.J. Flory, J. Am. Chem. Soc. 58 (1936) 1877. [69] C.P. Nicholas, W.E. Rathbun, T.M. Kruse, H.A. Pham, Composition of Oligomerate,
[21] G.J.P. Britovsek, R. Malinowski, D.S. McGuinness, J.D. Nobbs, A.K. Tomov, to UOP LLC, US Patent 9,644,159, May 9, 2017.
A.W. Wadsley, C.T. Young, ACS Catal. 5 (2015) 6922–6925. [70] W. van der Merwe, Environ. Sci. Technol. 44 (2010) 1806–1812.
[22] A.N. Parvulescu, D. Mores, E. Stavitski, C.M. Teodorescu, P.C.A. Bruijnincx, [71] S.D. Alexandratos, Ind. Eng. Chem. Res. 48 (2009) 388–398.
R.J.M. Klein Gebbink, B.M. Weckhuysen, J. Am. Chem. Soc. 132 (2010) [72] K.A. Mauritz, R.B. Moore, Chem. Rev. 104 (2004) 4535–4585.
10429–10439. [73] B.M. Antunes, A.E. Rodrigues, Z. Lin, I. Portugal, C.M. Silva, Fuel Process. Technol.
[23] 2015 Worldwide Refining Survey, Oil Gas J. (December) (2014). 138 (2015) 86–99.
[24] M.A. Hashim, J. Kulandai, R.S. Hassan, J. Chem. Technol. Biotechnol. 54 (1992) [74] S. Talwalkar, M. Chauhan, P. Aghalayam, Z. Qi, K. Sundmacher, S. Mahajani, Ind.
207–214. Eng. Chem. Res. 45 (2006) 1312–1323.
[25] Nonylphenol (NP) and Nonylphenol Ethoxylates (NPEs) Action Plan, US [75] M.L. Honkela, A.O.I. Krause, Catal. Lett. 87 (2003) 113–119.
Environmental Protection Agency, RIN 2070-ZA09, 2010. [76] R. Birkhoff, M. Nurminen, Handbook of Petroleum Refining Processes, in:

94
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

R.A. Myers (Ed.), 2004 1.3–1.9. E. Schmidt, G.W. Huber, J. Catal. 323 (2015) 33–44.
[77] J.D. Taylor, M.M. Jenni, M.W. Peters, Top. Catal. 53 (2010) 1224–1230. [130] J.W. Yoon, J.H. Lee, J.-S. Chang, D.H. Choo, S.J. Lee, S.H. Jhung, Catal. Commun.
[78] P.R. Gruber, M.W. Peters, J.M. Griffith, Y.A. Obaidi, L.E. Manzer, J.D. Taylor, D.E. 8 (2007) 967–970.
Henton, Renewable Compositions, to Gevo, Inc., US Patent 8,193,402, June 5, [131] K.H. Keuchler, S.H. Brown, A.A. Verberckmoes, S.E. Silverberg, M.P. Puttemans,
2012. M.R. Welford, J.S. Godsmark, Olefin oligomerization and compositions therefrom,
[79] M.W. Peters, J.D. Taylor, Renewable Jet Fuel Blendstock from Isobutanol, to Gevo, to ExxonMobil Chemical, US Patent 8,481,796, July 9, 2013.
Inc., US Patent 8,373,012, February 12, 2013. [132] J.A. Martens, R. Ravishankar, I.E. Mishin, P.A. Jacobs, Angew. Chem. Int. Ed. 39
[80] F.E. Shepard, J.J. Rooney, C. Kemball, J. Catal. 1 (1962) 379–388. (2000) 4376–4379.
[81] A. de Klerk, Energy Fuels 20 (2006) 1799–1805. [133] M.E. Loescher, T. Odegard, Oligomerization process, to Catalytic Distillation
[82] M. Misono, Y. Yoneda, Bull. Chem. Soc. Jpn. 40 (1967) 42–49. Technologies, US Patent 8,853,483, October 7, 2014.
[83] A. de Klerk, Energy Fuels 21 (2007) 625–632. [134] M.E. Loescher, D.G. Woods, M.J. Keenan, S.E. Silverberg, P.W. Allen,
[84] C.P. Nicholas, A. Bhattacharyya, D.E. Mackowiak, Process for Oligomerizing Oligomerization Process, to Catalytic Distillation Technologies and ExxonMobil
Dilute Ethylene, to UOP LLC, US Patent 8,748,682, June 10, 2014. Chemical, US Patent 7,145,049, December 5, 2006.
[85] R. Van Grieken, J.M. Escola, J. Moreno, R. Rodriguez, Appl. Catal. A 337 (2008) [135] M.M. Mertens, A. Verberckmoes, I.D. Johnson, Hydrocarbon conversion process
173–183. using molecular sieve of MFS framework type, to ExxonMobil Chemical, US Patent
[86] M. Golombok, J. de Bruijn, Ind. Eng. Chem. Res. 39 (2000) 267–271. 8,859,836, October 14, 2014.
[87] M. Golombok, J. de Bruijn, Appl. Catal. A 208 (2001) 47–53. [136] W.J. Roth, D.L. Dorset, G.J. Kennedy, Microporous Mesoporous Mater. 142 (2011)
[88] R. van Grieken, J.M. Escola, J. Moreno, R. Rodriguez, Chem. Eng. J. 155 (2009) 168–177.
442–450. [137] J.P.G. Pater, P.A. Jacobs, J.A. Martens, J. Catal. 184 (1999) 262–267.
[89] R. Van Grieken, J.M. Escola, J. Moreno, R. Rodriguez, Appl. Catal. A 305 (2006) [138] M. Bjorgen, K.-P. Lillerud, U. Olysbye, S. Bordiga, A. Zecchina, J. Phys. Chem. B
176–188. 108 (2004) 7862–7870.
[90] J.M. Escola, R. Van Grieken, J. Moreno, R. Rodriguez, Ind. Eng. Chem. Res. 45 [139] G.K. Chitnis, A.B. Dandekar, B.S. Umansky, G.B. Brignac, J. Stokes, W.A. Leet,
(2006) 7409–7414. National Petroleum Refiners Association, 2005 National Meeting, Paper AM-05-77
[91] M.R. Agliullin, I.G. Danilova, A.V. Faizullin, S.V. Amarantov, S.V. Bubennov, (2017).
T.R. Prosochkina, N.G. Grigor’eva, E.A. Paukshtis, B.I. Kutepov, Microporous [140] C.P. Nicholas, L.T. Nemeth, D.-Y. Jan, Olefin upgrading process, to Honeywell
Mesoporous Mater. 230 (2016) 118–127. UOP, US Patent 8,178,740, May 15, 2012.
[92] C. Flego, S. Peratello, C. Perego, L.M.F. Sabatino, G. Bellussi, U. Romano, J. Mol. [141] S.J. Frey, G.W. Fichtl, P. Barger, S.M. Roney, S.L. Krupa, C.P. Nicholas, Methods
Catal. A 204–205 (2003) 581–589. for Producing Jet-Range Hydrocarbons, US Patent Application 2015/0045599,
[93] D.H. Park, S.-S. Kim, T.J. Pinnavaia, F. Tzompantzi, J. Prince, J.S. Valente, J. Phys. February 12, 2015.
Chem. C 115 (2011) 5809–5816. [142] C.P. Nicholas, L.T. Nemeth, W. Plencner, C.L. Nicholas, T. Mezza, S. Prabhakar,
[94] C. Chizallet, P. Raybaud, Angew. Chem. Int. Ed. 48 (2009) 2891–2893. W. Sinkler, Appl. Catal. A 536 (2017) 75–84.
[95] F. Leydier, C. Chizallet, A. Chaumonnot, M. Digne, E. Soyer, A.-A. Quoineaud, [143] C.P. Nicholas, S.L. Krupa, K.M. Vanden Bussche, T.M. Kruse, Process for making
D. Costa, P. Raybaud, J. Catal. 284 (2011) 215–229. diesel by oligomerization, to Honeywell UOP, US Patent 9,441,173, September 13,
[96] F. Leydier, C. Chizallet, D. Costa, P. Raybaud, J. Catal. 325 (2015) 35–47. 2016.
[97] C. Chizallet, P. Raybaud, ChemPhysChem 11 (2010) 105–108. [144] G.W. Fichtl, C.P. Luebke, C.P. Nicholas, D.K. Sullivan, Process for the Production
[98] A.V. Lavrenov, V.K. Duplyakin, Kinet. Catal. 50 (2009) 235–240. of Jet-Range Hydrocarbons, to UOP, US Patent Application US2016/0312133,
[99] C. Perego, A. Carati, Zeolites and zeolite-like materials in industrial catalysis, October 27, 2016.
Chapter 14, Zeolites: From Model Materials to Industrial Catalysts, in: J. Čejka, [145] P.A. Jacobs, J.A. Martens, Stud. Surf. Sci. Catal. 33 (1987) 233–249.
J. Peréz-Pariente, W.J. Roth (Eds.), TransWorld Research Network, 2008. [146] P.A. Jacobs, J.A. Martens, Stud. Surf. Sci. Catal. 33 (1987) 251–274.
[100] C. Knottenbelt, J.P.P. Ntshabele, 18th World Petroleum Congress, Block 2, Forum [147] J.A. Martens, W.H. Verrelst, G.M. Mathys, S.H. Brown, P.A. Jacobs, Angew. Chem.
6, 2005. Int. Ed. 44 (2005) 5687–5690.
[101] M. Moliner, C. Martinez, A. Corma, Angew. Chem. Int. Ed. 54 (2015) 2–22. [148] G.M.K. Mathys, L.R.M. Martens, M.A. Baes, J.P. Verduijn, D.R.C. Huybrechts,
[102] C. Martinez, A. Corma, Coord. Chem. Rev. 255 (2011) 1558–1580. Alkene Oligomerization, to Exxon Chemical Patents, US Patent 5,672,800,
[103] http://www.iza-structure.org/databases/. September 30, 1997.
[104] M.B. Park, Y. Lee, A. Zheng, F.-S. Xiao, C.P. Nicholas, G.J. Lewis, S.B. Hong, J. Am. [149] L.R. Martens, J.P. Verduijn, G.M. Mathys, Catal. Today 36 (1997) 451–460.
Chem. Soc. 135 (2013) 2248–2255. [150] G.M.K. Mathys, S.H. Brown, H.J. Beckers, R.F. Caers, J.S. Godsmark, L.R.M.
[105] G.T. Kerr, Inorg. Chem. 5 (1966) 1537–1539. Martens, J.R. Shutt, E.T. Van Driessche, Olefin Oligomerization, to ExxonMobil
[106] R.J. Quann, L.A. Green, S.A. Tabak, F.J. Krambeck, Ind. Eng. Chem. Res. 27 (1988) Chemical, US Patent 6,884,914, April 26, 2005.
565–570. [151] C. Martinez, E.J. Doskocil, A. Corma, Top. Catal. 57 (2014) 668–682.
[107] X. Zhao, R.D. McGihon, S.A. Tabak, Hydrocarbon Eng. 9 (2008) 39–45. [152] S.H. Brown, J.S. Godsmark, G.M.K. Mathys, Olefin oligomerization process, to
[108] C. Knottenbelt, Catal. Today 71 (2002) 437. ExxonMobil Chemical, US Patent 8,716,542, May 6, 2014.
[109] K.H. Keuchler, S.H. Brown, H. Jaensch, G.M. Mathys, S. Luo, J.C. Cheng, Olefin [153] W.H. Verrelst, L.R.M. Martens, Oligomerization and catalysts therefore, to Exxon
Oligomerization to Produce Hydrocarbon Compositions Useful as Fuel, to Chemical, US Patent 6,143,942, November 7, 2000.
ExxonMobil Chemical, US Patent 7,678,954, March 16, 2010. [154] J.E. Stanat, G.M.K. Mathys, D.W. Turner, J.C. Cheng, S.W. Beadle, C.M.C.
[110] P. Borges, R.R. Pinto, M.A.N.D.A. Lemos, F. Lemos, J.C. Vedrine, E.G. Derouane, Guajardo, R. Eijkhoudt, A.D. Godwin, E.E. Green, C.M. Yarbrough, R.F. Caers, C.B.
F.R. Ribeiro, Appl. Catal. A 324 (2007) 20–29. Duncan, R.Y. Saleh, Oligomerization of Olefins, to ExxonMobil Chemical, US
[111] G. Bellussi, F. Mizia, V. Calemma, P. Pollesel, R. Millini, Microporous Mesoporous Patent 7,425,662, September 16, 2008.
Mater. 164 (2012) 127–134. [155] M. Kojima, M.W. Rautenbach, C.T. O’Connor, Ind. Eng. Chem. Res. 27 (1988)
[112] A. de Klerk, Energy Fuels 21 (2007) 3084–3089. 248–252.
[113] C.P. Nicholas, T.M. Kruse, Process for making gasoline by oligomerization, to [156] L.M. Tiako Ngandjui, F.C. Thyrion, Ind. Eng. Chem. Res. 35 (1996) 1269–1274.
Honeywell UOP, US Patent 9,278,893, March 8, 2016. [157] A.P. Vogel, C.T. O’Connor, M. Kojima, Clay Miner. 25 (1990) 355–362.
[114] C.M. Halmenschlager, M. Brar, I.T. Apan, A. de Klerk, Ind. Eng. Chem. Res. 55 [158] J.P.G. Pater, P.A. Jacobs, J.A. Martens, J. Catal. 179 (1998) 477–482.
(2016) 13020–13031. [159] G.D. Yadav, N.S. Doshi, Green Chem. 4 (2002) 528–540.
[115] S. Schwarz, M. Kojima, C.T. O’Connor, Appl. Catal. 68 (1991) 81–96. [160] M.L. Occelli, J.T. Hsu, L.G. Galya, J. Mol. Catal. 33 (1985) 371–389.
[116] A.G. Popov, V.S. Pavlov, I.I. Ivanova, J. Catal. 335 (2016) 155–164. [161] A. Mantilla, F. Tzompantzi, G. Ferrat, A. López-Ortega, E. Romero, E. Ortiz-Islas,
[117] A. Corma, C. Martinez, E. Doskocil, J. Catal. 300 (2013) 183–196. R. Gómez, M. Torres, Chem. Commun. (2004) 1498–1499.
[118] K.G. Wilshier, P. Smart, R. Western, T. Mole, T. Behrsing, Appl. Catal. 31 (1987) [162] J.S. Lee, J.W. Yoon, S.B. Halligudi, J.-S. Chang, S.H. Jhung, Appl. Catal. A 366
339–359. (2009) 299–303.
[119] M. Yamamura, K. Chaki, T. Wakatsuki, H. Okado, K. Fujimoto, Zeolites 14 (1994) [163] A. Mantilla, G. Ferrat, A. López-Ortega, E. Romero, F. Tzompantzi, M. Torres,
643–650. E. Ortiz-Islas, R. Gómez, J. Mol. Catal. A 228 (2005) 333–338.
[120] M. Henry, M. Bulut, W. Vermandel, B. Sels, P. Jacobs, D. Minoux, N. Nesterenko, [164] S.H. Jhung, J.-S. Chang, Catal. Surv. Asia 13 (2009) 229–236.
S. Van Donk, J.P. Dath, Appl. Catal. A 413–414 (2012) 62–77. [165] M.N. Timofeeva, Appl. Catal. A. 256 (2003) 19–35.
[121] A.N. Mlinar, P.M. Zimmerman, F.E. Celik, M. Head-Gordon, A.T. Bell, J. Catal. 288 [166] J.S. Vaughn, C.T. O’Connor, J.C.Q. Fletcher, J. Catal. 147 (1994) 441–454.
(2012) 65–73. [167] J. Zhang, R. Ohnishi, T. Okuhara, Y. Kamiya, Appl. Catal. A 353 (2009) 68–73.
[122] C.S.H. Chen, R.F. Bridger, J. Catal. 161 (1996) 687–693. [168] S. Yang, Z. Liu, X. Meng, C. Xu, Energy Fuels 23 (2009) 70–73.
[123] Z.-B. Yu, Y. Han, L. Zhao, S. Huang, Q.-Y. Zheng, S. Lin, A. Córdova, X. Zou, J. Sun, [169] Y. Gu, F. Shi, Y. Deng, Catal. Commun. 4 (2003) 597–601.
Chem. Mater. 24 (2012) 3701–3706. [170] O. Stenzel, R. Brüll, U.M. Wahner, R.D. Sanderson, H.G. Raubenheimer, J. Mol.
[124] M.J. Wulfers, R.F. Lobo, Appl. Catal. A 505 (2015) 394–401. Catal. A 192 (2003) 217–222.
[125] J.W. Yoon, J.-S. Chang, H.-D. Lee, T.-J. Kim, S.H. Jhung, J. Catal. 245 (2007) [171] B.L. Small, Acc. Chem. Res. 48 (2015) 2599–2611.
253–256. [172] Nexant Inc. ChemSystems Process Evaluation Research Planning (PERP)
[126] J.W. Yoon, S.H. Jhung, D.H. Choo, S.J. Lee, K.-Y. Lee, J.-S. Chang, Appl. Catal. A Report—Linear Alpha Olefins 2015-2, October 2015.
337 (2008) 73–77. [173] E.J. Arlman, P. Cossee, J. Catal. 3 (1964) 99–104.
[127] R. Schmidt, M.B. Welch, B.B. Randolph, Energy Fuels 22 (2008) 1148–1155. [174] S.A. Svejda, M. Brookhart, Organometallics 18 (1999) 65–74.
[128] A. Kulkarni, A. Kumar, A.S. Goldman, F.E. Celik, Catal. Commun. 75 (2016) [175] Y.V. Kissin, T.E. Nowlin, R.I. Mink, Macromolecules 26 (1993) 2151–2158.
98–102. [176] K. Takeuchi, T. Tamura, H. Tashiro, Process of Producing alpha-olefin, to Idemitsu
[129] Y.T. Kim, J.P. Chada, Z. Xu, Y.J. Pagan-Torres, D.C. Rosenfeld, W.L. Winniford, Petrochemical, US Patent 5,498,735, March 12, 1996.

95
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

[177] A.K. Tomov, J.D. Nobbs, J.J. Chirinos, P.K. Saini, R. Malinowski, S.K.Y. Ho, [225] E. Mazoyer, K.C. Szeto, N. Merle, J. Thivolle-Cazat, O. Boyron, J.-M. Basset,
C.T. Young, D.S. McGuinness, A.J.P. White, M.R.J. Elsegood, G.J.P. Britovsek, C.P. Nicholas, M. Taoufik, J. Mol. Catal. A 385 (2014) 125–132.
Organometallics 36 (2017) 510–522. [226] Y. Chen, E. Callens, E. Abou-Hamad, N. Merle, A.J.P. White, M. Taoufik,
[178] H.B. Fernald, R.G. Hay, A.N. Kresge, Conversion of Ethylene to Alpha Olefins in C. Coperet, E. Le Roux, J.-M. Basset, Angew. Chem. Int. Ed. 51 (2012) 11886.
the Presence of a Solvent, to Gulf Research and Development Corporation, US [227] Y. Chen, R. Credendino, E. Callens, M. Atiqullah, M.A. Al-Harthi, L. Cavallo, J.-
Patent 3,510,539, May 5, 1970. M. Basset, ACS Catal. 3 (2013) 1360–1364.
[179] H.B. Fernald, W. Gall, B.H. Gwynn, E.E. Nelson, Minimizing polymer formation in [228] Z. Xu, J.P. Chada, D. Zhao, C.A. Carrero, Y.T. Kim, D.C. Rosenfeld, J.L. Rogers,
the conversion of ethylene to a-olefins, to Gulf Research and Development S.J. Rozeveld, I. Hermans, G.W. Huber, ACS Catal. 6 (2016) 3815–3825.
Corporation, US Patent 3,502,741, March 24, 1970. [229] R.G. Schultz, J.M. Schuck, B.S. Wildi, J. Catal. 6 (1966) 385–396.
[180] C.W. Lanier, Process for the Production of Olefins, to Ethyl Corp, US Patent 3,789, [230] R.G. Schultz, R.M. Engelbrecht, R.N. Moore, L.T. Wolford, J. Catal. 6 (1966)
081, January 29, 1974. 419–424.
[181] E. Mazoyer, J. Trebosc, A. Baudouin, O. Boyron, J. Pelletier, J.-M. Basset, [231] R.G. Schultz, J. Catal. 7 (1967) 286–290.
M.J. Vitorino, C.P. Nicholas, R.M. Gauvin, M. Taoufik, L. Delavoye, Angew. Chem. [232] W. Zhang, W.-H. Sun, C. Redshaw, Dalton Trans. 42 (2013) 8988–8997.
Int. Ed. 49 (2010) 9854–9858. [233] R. Cariou, J.W. Shabaker, ACS Catal. 5 (2015) 4363–4367.
[182] B. Werghi, A. Bendjeriou-Sedjerari, A. Jedidi, E. Abou-Hamad, L. Cavallo, J.- [234] P.W.N.M. van Leeuwen, N.D. Clement, M.J.-L. Tschan, Coord. Chem. Rev. 255
M. Basset, Organometallics 35 (2016) 3288–3294. (2011) 1499–1517.
[183] A.H. Turner, J. Am. Oil Chem. Soc. 60 (1983) 623–627. [235] K.P. Bryliakov, E.P. Talsi, Coord. Chem. Rev. 256 (2012) 2994–3007.
[184] W. Keim, Angew. Chem. Int. Ed. Engl. 29 (1990) 235–244. [236] A.K. Tomov, V.C. Gibson, G.J.P. Britovsek, R.J. Long, M. van Meurs, D.J. Jones,
[185] W. Keim, Angew. Chem. Int. Ed. 52 (2013) 12492–12496. K.P. Tellmann, J.J. Chirinos, Organometallics 28 (2009) 7033–7040.
[186] P. Kuhn, D. Semeril, D. Matt, M.J. Chetcuti, P. Lutz, Dalton Trans. (2007) [237] R.Y. Brogaard, U. Olsbye, ACS Catal. 6 (2016) 1205–1214.
515–528. [238] D.S. McGuinness, Chem. Rev. 111 (2011) 2321–2341.
[187] F. Spieser, P. Braunstein, L. Saussine, Acc. Chem. Res. 38 (2005) 784–793. [239] T. Agapie, Coord. Chem. Rev. 255 (2011) 861–880.
[188] A. Finiels, F. Fajula, V. Hulea, Catal. Sci. Technol. 4 (2014) 2412–2426. [240] R.M. Manyik, W.E. Walker, T.P. Wilson, J. Catal. 47 (1977) 197–209.
[189] M.A. Deimund, J. Labinger, M.E. Davis, ACS Catal. 4 (2014) 4189–4195. [241] W.K. Reagan, B.K. Conroy, Chromium compounds and uses thereof, to Phillips
[190] A.N. Mlinar, G.B. Baur, G.G. Bong, A.B. Getsoian, A.T. Bell, J. Catal. 296 (2012) Petroleum Company, US Patent 5,288,823, February 22, 1994.
156–164. [242] W.K. Reagen, T.M. Pettijohn, J.W. Freeman, E.A. Benham, Process for olefin
[191] M. Tanaka, A. Itadani, Y. Kuroda, M. Iwamoto, J. Phys. Chem. C 116 (2012) polymerization, to Phillips Petroleum Company, US Patent 5,786,431, July 28,
5664–5672. 1998.
[192] C.P. Nicolaides, M.S. Scurrell, P.M. Semano, Appl. Catal. A 245 (2003) 45–53. [243] Y. Yang, Z. Liu, R. Cheng, X. He, B. Liu, Organometallics 33 (2014) 2599–2607.
[193] J.J. da Silva Ferreira Alves, C.P. Nicholas, Process for Oligomerizing Dilute [244] W.J. van Rensburg, C. Grove, J.P. Steynberg, K.B. Stark, J.J. Huyser,
Ethylene, to UOP LLC, US Patent Application 2011/0245567, October 6, 2011. P.J. Steynberg, Organometallics 23 (2004) 1207–1222.
[194] J.J. da Silva Ferreira Alves, J.E. Rekoske, C.P. Nicholas, Apparatus for Increasing [245] S. Tang, Z. Liu, X. Yan, N. Li, R. Cheng, X. He, B. Liu, Appl. Catal. A 481 (2014)
Weight of Olefins, to UOP LLC, US Patent 8,128,879, March 6, 2012. 39–48.
[195] H. Kurokawa, R. Ogawa, K. Yamamoto, T. Sakuragi, M.-A. Ohshima, H. Miura, J. [246] T.M. Zilbershtein, V.A. Kardash, V.V. Suvorova, A.K. Golovko, Appl. Catal. A 475
Jpn. Petrol. Inst. 57 (2014) 146–154. (2014) 371–378.
[196] E.D. Metzger, C.K. Brozek, R.J. Comito, M. Dinca, ACS Cent. Sci. 2 (2016) [247] J.Y. Jeon, D.S. Park, D.H. Lee, S.C. Eo, S.Y. Park, M.S. Jeong, Y.Y. Kang, J. Lee,
148–153. B.Y. Lee, Dalton Trans. 44 (2015) 11004–11012.
[197] E.D. Metzger, R.J. Comito, C.H. Hendon, M. Dinca, J. Am. Chem. Soc. 139 (2017) [248] I. Vidyaratne, G.B. Nikiforov, S.I. Gorelsky, S. Gambarotta, R. Duchateau,
757–762. I. Korobkov, Angew. Chem. Int. Ed. 48 (2009) 6552–6556.
[198] N. Ajellal, M.C.A. Kuhn, A.D.G. Boff, M. Horner, C.M. Thomas, J.-F. Carpentier, [249] D.S. McGuinness, P. Wasserscheid, W. Keim, C. Hu, U. Englert, J.T. Dixon,
O.L. Casagrande, Organometallics 25 (2006) 1213–1216. C. Grove, Chem. Commun. (2003) 334–335.
[199] J. Canivet, S. Aguado, Y. Schuurman, D. Farrusseng, J. Am. Chem. Soc. 135 (2013) [250] G.J.P. Britovsek, D.S. McGuinness, Chem. Eur. J. 22 (2016) 16891–16896.
4195–4198. [251] S. Heinig, A. Woehl, W. Mueller, M.H. Al-Hazmi, B.H. Mueller, N. Peulecke,
[200] Z. Li, N.M. Schweitzer, A.B. League, V. Bernales, A.W. Peters, A. Getsoian, U. Rosenthal, ChemCatChem 6 (2014) 514–521.
T.C. Wang, J.T. Miller, A. Vjunov, J.L. Fulton, J.A. Lercher, C.J. Cramer, [252] M. Gong, Z. Liu, Y. Li, Y. Ma, Q. Sun, J. Zhang, B. Liu, Organometallics 35 (2016)
L. Gagliardi, J.T. Hupp, O.K. Farha, J. Am. Chem. Soc. 138 (2016) 1977–1982. 972–981.
[201] H. Makio, T. Fujita, Acc. Chem. Res. 42 (2009) 1532–1544. [253] N. Cloete, H.G. Visser, I. Engelbrecht, M.J. Overett, W.F. Gabrielli, A. Roodt, Inorg.
[202] M. Watanabe, M. Kuramoto, Process for producing propylene oligomers, to Chem. 52 (2013) 2268–2270.
Idemitsu Kosan, US Patent 4,814,540, March 21, 1989. [254] A. Bollmann, K. Blann, J.T. Dixon, F.M. Hess, E. Killian, H. Maumela,
[203] S. Tanaka, Y. Shiraki, T. Tamura, M. Kuramoto, H. Sato, M. Watanabe, Catalysts D.S. McGuinness, D.H. Morgan, A. Neveling, S. Otto, M. Overett, A.M.Z. Slawin,
for alpha-olefin production and processes for producing alpha-olefin, to Idemitsu P. Wasserscheid, S. Kuhlmann, J. Am. Chem. Soc. 126 (2004) 14712–14713.
Petrochemical, US Patent 6,555,633, April 29, 2003. [255] D.S. McGuinness, P. Wasserscheid, W. Keim, D. Morgan, J.T. Dixon, A. Bollmann,
[204] A. Meiswinkel, A. Woehl, W. Mueller, H.V. Boelt, F.M. Mosa, M.H. Al-Hazmi, Oil H. Maumela, F. Hess, U. Englert, J. Am. Chem. Soc. 125 (2003) 5272–5273.
Gas J. 38 (2012) 103–106. [256] A.C. Pinheiro, T. Roisnel, E. Kirillov, J.-F. Carpentier, O.L. Casagrande Jr., Dalton
[205] A. Forestiere, H. Olivier-Bourbigou, L. Saussine, Oil Gas Sci. Technol. 64 (2009) Trans. 44 (2015) 16073–16080.
649–667. [257] T. Simler, P. Braunstein, A.A. Danopolous, Organometallics 35 (2016) 4044–4049.
[206] B. Hessen, H. van der Heijden, J. Am. Chem. Soc. 118 (1996) 11670–11671. [258] P. Ai, A.A. Danopoulos, P. Braunstein, Organometallics 34 (2015) 4109–4116.
[207] J.J.W. Eshuis, Y.Y. Tan, A. Meetsma, J.H. Teuben, J. Renkema, G.G. Evens, [259] I. Thapa, S. Gambarotta, I. Korobkov, M. Murugesu, P. Budzelaar, Organometallics
Organometallics 11 (1992) 362–369. 31 (2012) 486–494.
[208] K.C. Szeto, N. Merle, C. Rios, P. Rouge, J.L. Castelbou, M. Taoufik, Catal. Sci. [260] L. Zhang, X. Meng, Y. Chen, C. Cao, T. Jiang, ChemCatChem 9 (2017) 76–79.
Technol. 5 (2015) 4765–4771. [261] W.H. Monillas, J.F. Young, G.P.A. Yap, K.H. Theopold, Dalton Trans. 42 (2013)
[209] K. Ziegler, H. Martin, Production of dimers and low molecular polymerization 9198–9210.
products from ethylene, to Karl Ziegler, US Patent 2,943,125, June 28, 1960. [262] A.G.N. Coxon, R.D. Köhn, ACS Catal. 6 (2016) 3008–3016.
[210] S. Yamada, I. Ono, Bull. Japan Pet. Inst. 12 (1970) 160. [263] R.D. Köhn, M. Haufe, G. Kociok-Köhn, S. Grimm, P. Wasserscheid, W. Keim,
[211] J.A. Suttil, D.S. McGuinness, Organometallics 31 (2012) 7004–7010. Angew. Chem. Int. Ed. 39 (2000) 4337–4339.
[212] R. Robinson Jr., D.S. McGuinness, B.F. Yates, ACS Catal. 3 (2013) 3006–3015. [264] S. Liu, Y. Zhang, Y. Han, G. Feng, F. Gao, H. Wang, P. Qiu, Organometallics 36
[213] A. Sattler, J.A. Labinger, J.E. Bercaw, Organometallics 32 (2013) 6899–6902. (2017) 632–638.
[214] A. Sattler, D.C. Aluthge, J.R. Winkler, J.A. Labinger, J.E. Bercaw, ACS Catal. 6 [265] J. Heveling, C.P. Nicolaides, M.S. Scurrell, Appl. Catal. A 173 (1998) 1–9.
(2016) 19–22. [266] J. Andrews, P. Bonnifay, L.F. Albright, A.R. Goldsby (Eds.), The IFP Dimersol
[215] S. Dagorne, S. Bellemin-Laponnaz, C. Romain, Organometallics 32 (2013) Process for Dimerization of Propylene into Isohexenes: An Attractive Alternate to
2736–2743. Propylene Alkylation, Chapter 20, Industrial and Laboratory Alkylations, vol. 55,
[216] E. Despagnet-Ayoub, M.K. Takase, L.M. Henling, J.A. Labinger, J.E. Bercaw, 1977, pp. 328–340. ACS Symposium Series.
Organometallics 34 (2015) 4707–4716. [267] Y. Chauvin, A. Hennico, G. Leger, J.L. Nocca, Erdoel Erdgas Kohle 106 (7/8)
[217] R. Schmidt, M.B. Welch, R.D. Knudsen, S. Gottfried, H.G. Alt, J. Mol. Catal. A 222 (1990) 309–315.
(2004) 9–15. [268] N.A. Plechkova, K.R. Seddon, Chem. Soc. Rev. 37 (2008) 123–150.
[218] J.R.V. Lang, C.E. Denner, H.G. Alt, J. Mol. Catal. A 322 (2010) 45–49. [269] Y. Chauvin, S. Einloft, H. Olivier, Ind. Eng. Chem. Res. 34 (1995) 1149–1155.
[219] R. Schmidt, M.B. Welch, R.D. Knudsen, S. Gottfried, H.G. Alt, J. Mol. Catal. A 222 [270] J.M. Basset, Nature 519 (2015) 159.
(2004) 17–25. [271] Axens Dimersol G Process. http://www.axens.net/product/technology-licensing/
[220] S. Zhang, K. Nomura, J. Am. Chem. Soc. 132 (2010) 4960–4965. 10078/dimersol-g.html (Accessed 1 June 2017).
[221] R. Schmidt, P.K. Das, M.B. Welch, R.D. Knudsen, J. Mol. Catal. A 222 (2004) [272] Axens Dimersol X Process. http://www.axens.net/product/technology-licensing/
27–45. 10070/dimersol-x.html (Accessed 1 June 2017).
[222] K. Nomura, T. Mitsudome, A. Igarashi, G. Nagai, K. Tsutsumi, T. Ina, T. Omiya, [273] C.P. Nicolaides, M.S. Scurrell, P.M. Semano, Appl. Catal. A 245 (2003) 43–53.
H. Takaya, S. Yamazoe, Organometallics 36 (2017) 530–542. [274] J. Heveling, C.P. Nicolaides, M.S. Scurrell, Appl. Catal. A 248 (2003) 239–248.
[223] J. Romero, F. Carrillo-Hermosilla, A. Antiñolo, A. Otero, J. Mol. Catal. A 304 [275] C.P. Nicholas, A. Bhattacharyya, D.E. Mackowiak, Apparatus for Oligomerizing
(2009) 180–186. Dilute Ethylene, to UOP LLC, US Patent 8,021,620, September 20, 2011.
[224] C.P. Nicholas, T.J. Marks, Langmuir 20 (2004) 9456–9462 and references therein. [276] C.P. Nicholas, A. Bhattacharyya, D.E. Mackowiak, Process for Oligomerizing

96
C.P. Nicholas Applied Catalysis A, General 543 (2017) 82–97

Dilute Ethylene, to UOP LLC, US Patent 8,575,410, November 5, 2013. [299] M.C. Clark, B.S. Umansky, E.A. Nye, M.J. Reichensberger, W.C. Lewis, Process for
[277] C.P. Nicholas, A. Bhattacharyya, D.E. Mackowiak, Process for Oligomerizing making high octane gasoline with reduced benzene content by benzene alkylation
Dilute Ethylene, to UOP LLC, US Patent 8,748,681, June 10, 2014. at high benzene conversion, to ExxonMobil Research and Engineering, US Patent
[278] R.D. Andrei, M.I. Popa, C. Cammarano, V. Hulea, New J. Chem. 40 (2016) 8,395,006, March 12, 2013.
4146–4152. [300] E. Moy, S.C. Smyth, BenzOUT™ Technology for Benzene Reduction in Gasoline,
[279] M. Lallemand, A. Finiels, F. Fajula, V. Hulea, J. Phys. Chem. C 113 (2009) Chapter 1.3, Handbook of Petroleum Refining Processes, in: R.A. Meyers (Ed.), 4th
20360–20364. ed., McGraw Hill, 2016.
[280] W.E. Garwood, J.D. Kushnerick, S.A. Tabak, Catalytic process for oligomerizing [301] E.-M. El Malki, M. Clark, BenzOUT Reducing Benzene Enhancing Gasoline Product
ethene, to Mobil Oil, US Patent 4,717,782, January 5, 1988. Value, NPRA Conference, Phoenix, AZ, Mar. 21–23, 2010.
[281] M. Lallemand, O.A. Rusu, E. Dumitriu, A. Finiels, F. Fajula, V. Hulea, Stud. Surf. [302] S.-Y. H. Hwang, R. Birkoff, R.F. Guarino, J.E. Moy, J.C. Peters, Process for redu-
Sci. Catal. 174B (2008) 1139–1142. cing the benzene content of gasoline, to Badger Licensing, US Patent 9,200,215,
[282] M. Lallemand, O.A. Rusu, E. Dumitriu, A. Finiels, F. Fajula, V. Hulea, Appl. Catal. December 1, 2015.
A 338 (2008) 37–43. [303] T.M. Sakuneka, A. de Klerk, R.J.J. Nel, A.D. Pienaar, Ind. Eng. Chem. Res. 47
[283] A. Martinez, M.A. Arribas, P. Concepcion, S. Moussa, Appl. Catal. A 467 (2013) (2008) 1828–1834.
509–518. [304] B.G. Harvey, H.A. Meylemans, Green Chem. 16 (2014) 770–776.
[284] Y. Peng, M. Dong, X. Meng, B. Zong, J. Zhang, AIChE J. 55 (2009) 717–725. [305] M.E. Wright, B.G. Harvey, R.L. Quintana, Energy Fuels 22 (2008) 3299–3302.
[285] F. Tzompantzi, A. Mantilla, G. Del Angel, J.M. Padilla, J.L. Fernandez, J.A.I. Diaz- [306] M.E. Wright, B.G. Harvey, R.L. Quintana, Diesel and jet fuels based on the oli-
Gongora, R. Gomez, Catal. Today 143 (2009) 132–136. gomerization of butene, to the United States of America as Represented by the
[286] A. Mantilla, F. Tzompantzi, G. Ferrat, A. Lopez-Ortega, S. Alfaro, R. Gomez, Secretary of the Navy, US Patent 9,181,144, November 10, 2015.
M. Torres, Catal. Today 107–108 (2005) 707. [307] B.H. Babu, M. Lee, D.W. Hwang, Y. Kim, H.-J. Chae, Appl. Catal. A 530 (2017)
[287] V.N. Ipatieff, L. Schmerling, Ind. Eng. Chem. 38 (1946) 400–402. 48–55.
[288] A. de Klerk, R.J.J. Nel, Energy Fuels 22 (2008) 1449–1455. [308] D.M. Drab, H.D. Willauer, M.T. Olsen, R. Ananth, G.W. Mushrush, J.W. Baldwin,
[289] T.M. Saduneka, R.J.J. Nel, A. de Klerk, Ind. Eng. Chem. Res. 47 (2008) D.R. Hardy, F.W. Williams, Energy Fuels 27 (2013) 6348–6354.
7178–7183. [309] D.A. Simonetti, R.T. Carr, E. Iglesia, J. Catal 285 (2012) 19–30.
[290] C.P. Nicholas, A. Bhattacharyya, Apparatus for the reduction of gasoline benzene [310] M. Behl, J.A. Schaidle, E. Christensen, J.E. Hensley, Energy Fuels 29 (2015)
content by alkylation with dilute ethylene, to UOP LLC, US Patent 8,414,851, April 6078–6087.
9, 2013.
[291] C.P. Nicholas, A. Bhattacharyya, Apparatus for the reduction of gasoline benzene
content by alkylation with dilute ethylene, to UOP LLC, US Patent 8,747,785, June Christopher P. Nicholas is Principal Scientist at
10, 2014. Honeywell UOP. He joined Honeywell UOP in 2006 and has
[292] C.P. Nicholas, A. Bhattacharyya, Process for the reduction of gasoline benzene worked throughout the Research departments at UOP, pri-
content by alkylation with dilute ethylene, to UOP LLC, US Patent 8,895,793, marily focused on inventing and catalytically testing new
November 25, 2014. materials and processes. Particular foci have included het-
[293] S.H. Cha, J. Lee, J. Shin, S.B. Hong, Top. Catal. 58 (2015) 537–544. erogeneous catalytic processes such as olefin oligomeriza-
[294] G.C. Laredo, R. Quintana-Solorzano, J.J. Castillo, H. Armendariz-Herrera, tion and alkylation, synthesis of inorganic materials (in
J.L. Garcia-Gutierrez, Appl. Catal. A 454 (2013) 37–45. particular metal oxides and zeolites), process engineering,
[295] G.C. Laredo, J. Castillo, H. Armendariz-Herrera, Appl. Catal. A 384 (2010) molecular adsorption, and olefin metathesis. Chris is an
115–121. inventor or co-inventor on more than 75 US and foreign
[296] G.C. Laredo, J. Castillo, J.O. Marroquin, F. Hernandez, Appl. Catal. A 363 (2009) patents and coauthor of 20+ peer-reviewed journal articles
11–18. and a book chapter. Chris’ research interests span the
[297] J.C. Cheng, T.J. Huang, Process for the alkylation of benzene-rich reformate using breadth of inorganic and catalytic technologies ranging
MCM-49, to Mobil Oil, US Patent 5,545,788, August 13, 1996. from materials synthesis to characterization to catalyst and process development. He has
[298] B.S. Umansky, M.C. Clark, A.B. Dandekar, Gasoline production by olefin poly- also served in multiple roles for the Chicago section of the North American Catalysis
merization with aromatics alkylation, to ExxonMobil Research and Engineering, Society. Prior to joining UOP, Chris earned a B.A. from Kalamazoo College, and a Ph.D.
US Patent 7,525,002, April 28, 2009. from Northwestern University.

97

Vous aimerez peut-être aussi