You are on page 1of 16

Fuel xxx (xxxx) xxx

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Review article

Effect of metal–acid balance and textual modifications on


hydroisomerization catalysts for n-alkanes with different chain length: A
mini-review
Chenhao Wei, Guohao Zhang, Liang Zhao *, Jinsen Gao, Chunming Xu
The State Key Lab of Heavy Oil Processing, China University of Petroleum (Beijing), Beijing 102249, China

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

Keywords: Hydroisomerization of n-alkanes has become an essential route for producing high quality fossil fuels and lu­
Hydroisomerization bricants meeting the state-of-art regulations. The hydroisomerization of light alkanes, i.e., pentane and hexane,
Metal-acid balance has been well developed. However, this process for heavier n-alkanes still has a large room for improvement.
Mechanism
Therefore, the purpose of the present work is to review similarities and differences among the hydro­
Hierarchical structure
isomerization of n-alkanes with different chain length. Attention has been paid on the effect of metal–acid
N-alkane
balance and textual modifications on the bifunctional catalysts. Besides, the mechanism of hydroisomerization
was also discussed in detail. At last, conclusions and prospects for the hydroisomerization of long-chain n-alkanes
have been made. This review could help to understand difficulties in enhancing isomers yield during the
hydroisomerization process and could benefit new researchers in this field by relating the state-of-art researches
to the reaction mechanism.

1. Introduction alkanes hydroisomerization, nowadays, the most studied and promising


catalysts are zirconia-based ones and zeolites-based ones, such as Pt/
Hydroisomerization of normal alkanes have drawn great attention in SO42− -ZrO2, Pt/WO3-ZrO2 and Pt/MOR, Pt/BEA, Pt/Y [11–17]. Plat­
petroleum industry as well as in exploiting Fischer–Tropsch oil and bio- inum promoted sulfated zirconia and tungsten zirconia catalysts show
derived oil in recent years [1–3]. N-alkanes make up a great percentage great activity. The advantages of zeolites-based catalysts have great
in distillated crude oil cuts and Fischer–Tropsch oil. As is shown in tolerance on water and sulfur contaminants and they can be easily re­
Table 1, hydroisomerization of C4-C8 n-alkanes can greatly increase the generated. While for long-chain alkanes, the most used acid supports are
octane number of gasolines to compensate for the loss caused by regu­ zeolites and silicoaluminophosphate molecular sieves with unidirec­
lations of olefin contents, while the hydroisomerization of C10-C22 can tional pore system [18]. To date, Pt/SAPO-11 and Pt/ZSM-22 are the
benefit the low-temperature flow properties of diesel, for example, the most investigated ones [19,20]. The first advantage of using these uni­
cold filter plugging point, viscosity, and freezing point [4–7]. Especially directional molecular sieves is their unique shape selectivity for iso-
for the n-alkanes with carbon number greater than 20, which are usually alkanes, including reactant shape selectivity, transition state shape
called wax, through hydroisomerization one can produce lubricants selectivity and product shape selectivity [21]. The second advantage is
with ultra-high viscosity index and much lower pour point [8]. There­ their stability by reducing coke formation.
fore, the techniques of hydroisomerization is not only environment In practice, the development and research status of n-alkanes
friendly, but also able to gain economical profits. hydroisomerization with different chain length is unbalanced. For
Catalysts used for hydroisomerization are usually bifunctional, example, hydroisomerization of n-pentane and n-hexane has been
containing hydrogenation-dehydrogenation function and acid catalysis extensively studied, and there has been several industrial applications
function. The former one is usually supplied by noble metals, such as with isomers yield higher than 95% [19,22,23]. While there are still
platinum and palladium, concerning their high (de)hydrogenation ac­ limitations on the industrial application of hydroisomerization of n-
tivity. While the latter one is usually gifted by solid acids. For light heptane and n-octane, concerning the low yield of desired multi-

* Corresponding author.
E-mail address: liangzhao@cup.edu.cn (L. Zhao).

https://doi.org/10.1016/j.fuel.2021.122809
Received 28 September 2021; Received in revised form 20 November 2021; Accepted 1 December 2021
0016-2361/© 2021 Elsevier Ltd. All rights reserved.

Please cite this article as: Chenhao Wei, Fuel, https://doi.org/10.1016/j.fuel.2021.122809


C. Wei et al. Fuel xxx (xxxx) xxx

Table 1
The properties of different n-alkane molecules and their isomers. Data is adapted
from ref. [9,10].
Category Compound Properties Value

Gasoline n-heptane Research octane number 0


2-methylpentane 46
2,3-dimethylbutane 89
Diesel n-hexadecane Cetane number 100
n-hexadecane Melting point (◦ C) 17.9
3-methylpentadecane Cetane number 87
3-methylpentadecane Melting point (◦ C) − 21.2
Lubricants n-hexacosane Viscosity Index 188
n-hexacosane Pour point (◦ C) 56.2
11-butyldocosane Viscosity Index 128
11-butyldocosane Pour point (◦ C) 0

branched products. As for the hydroisomerization of n-alkanes with


ultra-long chain, which is called isodewaxing, to produce lubricants
meeting API III and API III+ viscosity index, the existing industrial ap­
plications are still facing the challenges of low yields and high energy
consumption.
In this review, we attempt to give comparative views on the hydro­ Scheme 1. Classic mechanism of n-alkane hydroisomerization over bifunc­
isomerization between linear light alkanes (n-pentane and n-hexane) tional catalysts.
and long-chain alkanes (no less than n-heptane). The first part focus on
the summary and comparison of hydroisomerization mechanism. The the less limitations on diffusion of small molecules. While for long-chain
effects of metal–acid balance were discussed in the second part. Then n-alkenes, it’s not easy for the whole molecule diffuse from one side of
recent development and modifications on the textual properties of cat­ the micropore to another, due to the diffusion constrain imposed by the
alysts were summarized for the hydroisomerization of long-chain al­ interaction between alkene molecule and the wall of micro channels. In
kanes. Through the summarized research and comparison, we hope such case, to explain how is the long-chain n-alkenes are isomerized, two
lights can be shed into the development of hydroisomerization catalysts unique reaction mechanisms were proposed, which are so called “pore
for long-chain alkanes. mouth” and “key lock” mechanism [31–34]. “Pore mouth” mechanism
describes the situation in which one end of the carbon chain of n-alkene
2. Mechanism of n-alkane hydroisomerization molecule penetrates into the micropores, and it is catalyzed by the acid
sites located near the pore mouth. This unique mechanism is usually
2.1. The overall mechanism of hydroisomerization used to explain the formation of mono-branched products whose
branching is near the terminal position. While “key lock” mechanism
Generally, the hydroisomerization reactions are based on bifunc­ describes the situation in which both two ends of the carbon chain
tional catalysts, which comprise hydrogenation-dehydrogenation func­ penetrate two adjacent pore openings simultaneously. The same as it in
tion and acid function. It is widely accepted that the acid sites are active “pore mouth” mechanism, the isomerization is catalyzed by the acid
in this process only if they are in their Brønsted form instead of Lewis sites near the pore mouth. This is used to explain the central branching
form [24–26]. The most used catalysts include Pt and Pd as of isomer products. For light alkanes, the “key lock” mode is not favored,
hydrogenation-dehydrogenation active sites and molecular sieves as because the penetration into two adjacent pore openings requires the
acid supports [27–29]. Thus, the following discussion is based on chain length being greater than the thickness of the pore wall. For long-
traditional bifunctional catalysts. chain alkanes, it’s said the “key lock” catalysis contributes significantly
The classic bifunctional mechanism of n-alkane hydroisomerization in the distribution of methyl-branched isomers. And this contribution
was proposed by Coonradt and Garwood [30], and is shown in Scheme increases with the carbon number of reactants. Furthermore, high re­
1. It contains the following 7 successive steps: action temperature was regarded in favor of “key lock” mechanism,
considering the adsorption entropy is lower than that in the “pore
a) Diffusion of n-alkane molecules from bulk phase to the metal sites. mouth” mode [31].
b) Dehydrogenation of n-alkane molecules on metal sites to form n- In the hydroisomerization of n-alkanes, the chain length has signif­
olefin intermediates. icant influence on the products distribution. Though the overall
c) Diffusion of n-olefin intermediates from the metal sites to Brønsted hydroisomerization mechanism of both light alkanes and long-chain
acid sites, which is either in the micropores or on the external surface alkanes is very similar, the yields of isomers are much lower in the
of molecular sieves. hydroisomerization of long-chain alkanes. For example, the yields of
d) Isomerization of n-olefin intermediates on the acid sites to form iso- isomer products in C5/C6 n-alkanes hydroisomerization industrial ap­
olefins. plications, such as Hyysomer HS-10 from UOP, Hysopar 5000 from
e) Diffusion of iso-olefins from the acid sites to metal sites. Clariant, have reached to more than 95%, which can be regarded suc­
f) Hydrogenation of iso-olefins on the metal sites to form iso-alkanes. cessful [23]. However, compared to light alkane isomerization, the
g) Diffusion of iso-alkanes to bulk phase. yields of isomerized products in the hydroisomerization of long-chain n-
alkanes are much lower. In academic studies, this number range from
Obviously, there are 3 chemical steps and 4 physical steps. And in about 30% to 70% [35–38]. The significant difference in isomers yields
these 3 chemical steps, the dehydrogenation and hydrogenation steps are mainly attributed to the following two reasons. The first is the
are much faster than the isomerization step. Once n-alkenes were different levels of diffusion limitations imposed by catalysis materials
absorbed on Brønsted acid sites, the isomerization may happen. For light [39]. For alkanes, the molecular dimensions roughly increase with
alkenes, the Brønsted acid sites inside the micropores of 10 or 12- carbon number, thus, the severity of spatial confinement imposed by
membered rings molecular sieves are usually accessible, considering shape selective zeolites also increases. This will lead to higher possibility

2
C. Wei et al. Fuel xxx (xxxx) xxx

of excessive reactions during the diffusion of reactants, and further lead


to cracking. The second reason is the difference on isomerization
mechanism of alkene intermediates with different chain length.
Compared to light alkanes (C5/C6), more skeletal cleavage pathways are
available for alkanes whose carbon number is equal to or greater than 7,
and the detailed mechanism will be discussed in the next subsection.

2.2. Isomerization of alkene intermediates

To get better understanding of hydroisomerization, the isomeriza­


tion mechanism of alkene intermediates shall be discussed in detail.
Isomerization of alkenes is based on carbenium ions transformation. The
carbenium ions are formed on Brønsted acid sites by protonation.
Alvarez et al. [40] has summarized the isomerization and cracking re­
actions of carbenium ions.
The isomerization reactions of carbenium ions was classified into Scheme 3. Isomerization of carbenium ions via a protonated cyclobutene
two kinds: Type A isomerization occurs through alkyl shift, in which the intermediate.
position of branching changes but the degree of branching does not
change. Type B isomerization is the originates of side chains. It occurs 2.3. Undesired cracking
through protonated cyclopropane (PCP) intermediates and protonated
cyclobutene (PCB) intermediates, as shown in Scheme 2 and Scheme 3. The isomerization of carbenium ions is always accompanied with
PCP intermediate explains the formation of mono-methyl- and multi- cracking reactions, and the cracking products are undesired at most time
methyl-isomers, which can be transformed from low branching degree because of their low value. As mentioned above, the isomer yields of
to high branching degree successively [41]. While the PCB intermediate light alkanes are much higher than that of long-chain alkanes in
is proposed to explain the formation of ethyl-branched isomers. Of note hydroisomerization. To explain and understand this phenomenon, the
is that, due to the effect of steric hindrance, PCB intermediate could detailed discussion of cracking mechanism is necessary.
rarely format inside 10-membered ring channels. Besides, from the en­ β-scission refers to the cleavage of carbon–carbon bond at the Beta
ergetic view, the formation of protonated cyclopropane intermediates position of charged carbenium ion. The cracking reactions, i.e. β-scis­
from classical secondary carbenium have a lower energy barrier than sion, happen through 4 types of mechanisms, which are called type A,
that of protonated cyclobutene intermediates [42]. This explains the type B, type C, and type D mechanisms. As shown in Scheme 4, these 4
phenomenon that in the hydroisomerization of n-alkanes, the observed types of cracking mechanisms are classified according to the different
ethyl-branched isomers are always rare compared with methyl- structures of carbenium ions participated. Type A pathway includes two
branched ones [10,43,44]. tertiary carbenium ions, it is only possible when the carbon number of
These two kinds of isomerization reactions, i.e., type A and type B, isomers is no less than 8. Type B (including Type B1 and Type B2)
are reported to have different relative reaction rates. In hydro­ pathway includes one secondary and one tertiary carbenium ions, either
isomerization of n-decane, the methyl shift reactions were reported to be from secondary to tertiary or from tertiary to secondary ones. This re­
faster than PCP branching in 70 times [45]. In n-hexane hydro­ quires the reactant no less than 7 carbon atoms. Type C pathway in­
isomerization, the transformation between mono-methylpentanes is cludes two secondary carbenium ions, the original chain length should
about 10 times faster than branching degree changes. And Alvarez et al. not be less than 6. The last Type D pathway includes the formation of
[40] used approach to thermodynamic equilibrium (ATE) to get the unstable primary carbenium ions from secondary ones. The rate con­
estimation of the relative rates of type A and type B isomerization. The stants of these 4 modes of cracking reactions keep decrease from type A
result shows, in n-decene hydroisomerization, methylnonenes formed to type D [46]. From the above cracking mechanisms, one can tell that in
through protonated cycloalkane intermediates is 6–7 times slower than the hydroisomerization of n-alkanes, light alkanes (C5/C6) have less
through methyl shift. Though difference on those relative rates exists, it chance of skeletal cleavage than long-chain alkanes do.
still can be concluded that type A isomerization is much faster than type In F. Alvarez et al.’s [40] research of hydroisomerization and hy­
B isomerization. drocracking of n-decane, the relative rates of isomerization and cracking
reactions were evaluated according to the distribution of isomers and
cracking products under the reaction condition of T = 473 K, pH2 = 0.9
bar, pn-decane = 0.1 bar. As shown in Scheme 5, the rate constants of
successive isomerization from n-C10 to trimethyl isomers are similar.
However, there are significant differences in rates of cracking reactions
between isomers with various branching degrees. Type C and type D
cracking reactions correspond to the reactants of n-decane and mono­
branched isomers. For type D cracking, the products contain the un­
stable primary carbenium ions, and the rate is very slow. While for type
C cracking, the reaction rates are too slow compared to the successive
isomerization and other cracking reactions to be taken into account
[47]. Therefore, the value of their relative rates is not shown in the
scheme. Most of the cracking products are formed through type A and
type B cracking. And as shown in the scheme, the cracking of tri­
branched isomers is about 50 times faster than the cracking of
dibranched isomers. Of note is that, all the cracking mechanism and
their relative rates are obtained on catalysts with zeolite Y or USY as the
Scheme 2. Isomerization of carbenium ions via a protonated cyclopropane acid supports, and these zeolites can be considered has negligible
intermediate.

3
C. Wei et al. Fuel xxx (xxxx) xxx

Scheme 4. Classifications of cracking reactions. Adapted from [46] with permission from Elsevier.

Scheme 5. Relative rate constants of isomerization and cracking. Adapted from [40] with permission from Elsevier.

diffusion limitations on the reactants and reaction intermediates 3.1. Comparison of the effects of nMe/nA value
because of their large pore size.
Through the above analysis of isomerization and cracking reactions To investigate the effects of metal–acid balance and to give a quan­
mechanism, the conclusion can be drawn as follows: titative evaluation, the nMe/nA value is the most commonly used
parameter. This value refers to the ratio of accessible metal sites to the
a) Compared to long-chain n-alkanes, light alkanes (C5/C6) can only accessible acid sites with given strength. For the purpose of obtaining
crack through type C and type D mechanism, which are much slower systematic investigation, various methods were used to tuning the
than the rate of isomerization. Therefore, it is very promising to get metal–acid balance. The most commonly used methods include tuning
high yields of isomers in pentane and hexane hydroisomerization. the Si/Al ratio of molecular sieves to change the Brønsted acid density/
b) From n-heptane, with the increase of chain length, the overall strength, and changing the metal loading of synthesized catalysts.
cracking reactions are extensively enhanced considering both the Alvarez et al. [48] prepared HY zeolites with Si/Al ratio of 3 and 9 by
intrinsic type A, type B cracking and the various structures of tri­ calcination and SiCl4 dealumination of USYNH4, respectively. By
branched isomers. This indicates the precise control of branching changing the Pt loading and Si/Al ratio, catalysts with nPt/nA from 0.002
degree is of greater importance for reactants with longer chain length to 0.167 were obtained. For n-hexane transformation, the effects of nPt/
to enhance the isomers yields. nA can be divided into three stages. The first stage refers to the low
platinum content, nPt/nA < 0.006. At this stage, the (de)hydrogenation
3. Effects of metal–acid balance and the comparison between function have not reached to the balance with acid function. The (de)
alkanes with different chain length hydrogenation reactions appeared to be the rate limiting step. Hexene
produced by Pt sites could hardly compensate for the consumption on
As is shown in the above reaction mechanisms, the performance of acid sites. Furthermore, it limited the in-time hydrogenation of iso­
hydroisomerization is determined by both the characteristics of (de) merized alkenes. Due to the relative excess of Brønsted acid sites, those
hydrogenation function and Brønsted acid function, i.e., metal–acid isomerized alkenes have a greater possibility to undergo further reac­
balance. Their balance is mainly determined by the ratio of accessible tion. As mentioned in the mechanism, this will lead to cracking as well as
metal sites to acid sites, which is referred as nMe/nA or CMe/CA, and the coke formation. Therefore, the activity and stability of catalysts were
properties of acid supports, including acid strength, acid density and limited, and kept increase with increasing nPt/nA. Mono-branched and
acid location. However, for light alkanes and long-chain alkanes, the di-branched isomers both appeared to be the primary products, and
effects of the balance are not totally the same. Detailed discussion was cracking products appeared even at low conversion. When 0.006 < nPt/
given in this chapter. nA < 0.1, it reached the second stage. The activity of the catalyst reached
a plateau. Moreover, only mono-branched isomers were the primary
products. However, the ratio of mono-branched to di-branched isomers

4
C. Wei et al. Fuel xxx (xxxx) xxx

in the products kept increasing from about 25 to 30 with increasing nPt/ Alvarez et al. [40] studied the effects of nPt/nA in the hydroisomerization
nA. It seems the olefinic intermediates produced by (de)hydrogenation of n-decane on a series of Pt/HY catalysts with nPt/nA from 0.002 to
sites were able to feed all the acid sites. However, in this nPt/nA range, 0.48. From the results, the effects of nPt/nA ratio still can be divided into
the stability of catalysts had not reached the maximum. It indicated three stages, with the range of less than 0.03, greater than 0.03 but less
though all the acid sites have participated in reactions, some of them than 0.17, and greater than 0.17. All the features in those 3 stages were
were still reacting with reactants that have undergone at least one the same as in ref. [50]. Batalha et al. [51] studied the effects of met­
isomerization. While for nPt/nA > 0.1, it can be regarded to the third al–acid balance and metal–acid intimacy for hydroisomerization of n-
stage, in which it preserved most features in the second stage. However, hexadecane using HBEA zeolites as the acid support. They made three
the ratio of mono-branched to di-branched isomers kept constant. No series of Pt loaded HBEA catalysts according to the intimacy difference.
deactivation caused by coke formation was observed in 3 h. The reaction The concentration of Brønsted acid sites were represented by CH+, which
scheme seems to be perfectly obeys the successive carbenium ion included those acid sites able to retain pyridine at 150 ◦ C. On these three
transformation introduced above. It means every olefinic intermediate series of catalysts, similar effects of CPt/CH+ ratio were found, and the
only undergoes one isomerization reaction during the diffusion from one main difference is the exact values of nodes between the first and the
metal site to another. And the isomerization reactions became the rate second stage. Those CPt/CH+ values were 0.0040, 0.015 and 0.015 for
limiting step. Therefore, at this stage, the catalysts were called “ideal” these three series of catalysts, respectively.
bifunctional catalysts. From the above results and discussion, one can get some insights
Similar results were obtained by Lyu et al. [49] on Ni loaded SAPO- about the effects of CMe/CA (or nMe/nA): 1) For hydroisomerization of
11 catalysts. They studied the hydroisomerization of n-hexane on the both light alkanes and long-chain alkanes, the effects of CMe/CA ratio can
catalysts with CNi/CA range from 0 to 0.30. The result showed the ac­ be divided into three stages. In these three stages, the activity, stability
tivity of the catalysts, which is represented by turnover frequency (TOF) of catalysts and products distribution varies. The best hydro­
per acid site, kept increasing at CNi/CA less than 0.19. At this range, both isomerization performance of catalysts can be obtained at the third
mono-branched (MP) and di-branched (DMP) isomers were primary stage. However, it is worth noting that even the high CMe/CA is necessary
products. For example, on the catalyst SN-1, whose CNi/CA = 0.10, the for high yield of isomers and maximal stability, over loaded metal can
DMB/MP value equals to 0.02 when it is extrapolated to zero conver­ also cause negative effects, such as hydrolysis reaction and shadowing
sion. These features have well met that of the first stage mentioned acid sites [52–55]. 2) The effects of CMe/CA ratio is not the same for light
above. While for CNi/CA > 0.19, DMB/MP ratio and cracking/isomeri­ alkanes and long-chain alkanes. The difference was concluded in the
zation ratio can be observed close to 0 at low conversion. This means the following two aspects: a) In the second stage, the cracking reaction for
hydroisomerization reaction starts to follow the consecutive scheme, light alkanes exists but not as obvious as it for long-chain alkanes. b) the
which is the same as observed from the second stage mentioned above. value of CMe/CA ratio is not the same. The first aspect can be attributed
Indeed, the ratio of CMe/CA values varied a lot in the comparison with it to the higher reactivity of long-chain alkanes [56,57]. While the second
in the last quoted research. However, it was caused by the (de)hydro­ aspect may not only due to the reactivity gap but also attributed to the
genation activity difference of metal Pt and Ni. In comparison, each different catalysts and reaction conditions studied.
platinum sites can feed about 170 acid sites with olefinic intermediates,
while this number is only 5 for nickel sites. 3.2. Comparison of the effects of acid function
For the hydroisomerization of long-chain alkanes, e.g., n-heptane,
the effects of nPt/nA were studied by Ginenetto et al. [50]. They used the The effects of acid function of the catalysts have an essential role in
same Pt-loaded HY zeolites with Si/Al ratio of 3 and 9 as ref. [48]. While hydroisomerization of normal alkanes, including acid strength, acid
in the original article, the effects of nPt/nA value were discussed in the density and spatial distribution of acid sites. However, those effects have
form of nA/nPt. To get a better comparison, they were converted into the remarkable difference between light alkanes and long-chain alkanes.
same form as others in this review. The same as it for light alkanes, the
effects of nPt/nA for n-heptane can also be divided into three stages. At 3.2.1. Effects of acid strength and density
nPt/nA < 0.029, all the isomers including monobranched, dibranched In the hydroisomerization of light alkanes on zeolites-based cata­
and tribranched ones were formed directly, and cracking products lysts, the tuning of acidity was usually along reducing the total acid sites
appeared at zero conversion. Till the conversion reached to 50%, the but increasing the density of strong acid sites.
distribution of cracking products had reached more than 90%. This For n-hexane hydroisomerization, Viswanadham et al. [58] studied
indicated both isomers and cracking products were primary products. the performance of dealuminated mordenite based catalysts. The parent
The lack of (de)hydrogenation function caused the excessive reaction of mordenite was dealuminated by steaming and nitric acid. Significant
each olefinic intermediates. Then the higher cracking tendency of over- loss of total acidity was observed after dealumination. However, the
isomerized products contributed to the cracking products and limited strong acid sites increased from 0.36 mmol/g in HM to 0.42 mmol/g in
the isomer selectivity. Of note is that the deactivation caused by coke S3-HM. Through the catalyst evaluation in a fixed bed, the conversion of
formation was observed. When nPt/nA came to between 0.029 and 0.10, n-hexane increased by 20% at the same reaction conditions, while the
from the products distribution, it seems only isomers were the primary isomers selectivity kept almost constant at close to 100%. Thus, the
products. And the activity of PtHY catalysts came to be constant. Though maximum yield of isomers was obtained on dealuminated samples.
cracking products were also obvious, from 41.4% to 20.7% with Similar results were obtained by Saxena et al. [59] in the hydro­
increasing nPt/nA at 50% conversion, the value was reduced to a large isomerization of n-hexane. In their research, mordenite was used as
extent. It indicated olefinic intermediates will encounter serval acid sites parent zeolites to prepare catalysts, denoted as HM. Dealumination of
and undergo more than one reaction before being hydrogenated. How­ mordenite was performed by acid leaching with increasing concentra­
ever, for most intermediates, those acid sites were not enough for the tion of nitric acid, the samples were denoted as HM1, HM2 and HM3
formation of cracking products. At the third stage, nPt/nA > 0.10, respectively. By acid leaching, the total acid density decreased following
monobranched, dibranched, tribranched isomers and cracking products the sequence HM > HM1 > HM2 > HM3. However, the acid density of
were observed being formed step by step. At 50% conversion of n-hep­ strong ones changed inversely. This further influenced the conversion
tane, the minimum distribution of cracking products was only 5.3%. and cracking selectivity on Pt-loaded catalysts. Compare to Pt/HM, the
Since each olefinic intermediates only went through one isomerization cracking selectivity on Pt/HM1 to Pt/HM3 was reduced from 47% to
before being hydrogenated, the formation of multi-branched isomers 2.5% maximum. The conversion was increased from 21.1% to 31.0%
with higher cracking tendency was in low possibility. minimum. Thus, the isomer yields were increased. The enhanced cata­
For alkanes with longer chain length, there are also similar research. lytic performance was attributed to the decrease of total acid density but

5
C. Wei et al. Fuel xxx (xxxx) xxx

increased strong acid sites. However, the interpretation of how the the cracking.
change of acidity influences the performance was still lacking, and it Liu et al. [65] reported the synthesis of Fe substituted ZSM-22. The
requires further study. Here, we try to give some conjectures. Firstly, the decrease of Brønsted acidity was obtained by Fe substitution in the
reactivity of linear light alkenes is not high enough for their reactions on framework of ZSM-22. Acidity analysis of Fe-containing ZSM-22 showed
medium and weak acid sites. Only strong ones can catalyze the initial that compared to conventional ZSM-22 zeolites, the addition of Fe had
isomerization. This explains the increased conversion on catalysts with evidently decreased the Brønsted acid density and strength. The results
more strong acid sites. Secondly, however, the successive isomerization of n-dodecane hydroisomerization showed higher selectivity was ob­
may have a higher possibility to happen on medium and weak acid sites. tained on Fe substituted catalysts. For example, the weak and strong
That’s why the isomers selectivity increased with the decreasing total Brønsted acid density of H-ZF-1 was reduced by 20 μmol g− 1 and 63
acid sites. μmol g− 1 compared to conventional H-ZSM-22, respectively. And the
While in the hydroisomerization of long-chain alkanes, the general isomers selectivity on Pt/H-ZF-1 was always higher than conventional
idea was to lower the acid strength and acid density. Chen et al. [60] Pt/H-ZSM-22 by about 25% at each conversion. The reduction of
reported a bundle-shaped ZSM-22 zeolites synthesis with modified acid Brønsted acid density and strength was attributed to framework Al was
strength. Compared to conventional hydrothermal synthesis, the syn­ partially substituted by Fe which has low polarizability. And Fe-OH-Si
thesis conditions were tuned by the addition of ethanol. They found the bridge hydroxyls inside the zeolite framework have lower Brønsted
difference of acid strength and Al atoms coordination environment be­ acid strength than Al-OH-Si bridge hydroxyls. The low acid strength and
tween these samples. Compared with conventional ZSM-22, the strong acid density were believed to have better performance on the desorption
acid sites of the bundle-shaped samples were reduced from 59.65 μmol of isomer products and can maintain better metal–acid balance, there­
g− 1 to 36.10 μmol g− 1, and the weak ones were reduced from 74.98 fore the isomers selectivity was enhanced.
μmol g− 1 to 49.99 μmol g− 1. It was regarded that the change of Al atom It seems for long-chain alkanes, modifications on the zeolites for
coordination environment was responsible for the decrease of Brønsted reducing the acid strength and density were efficient ways to enhance
acid strength. The result of n-dodecane hydroisomerization showed that the isomers selectivity. However, different opinion was given by Noh
the mild Brønsted acid strength of bundle-shaped ZSM-22 led to the et al. [66]. They studied the relation between acid strength and isom­
selectivity enhancement from 43.82% to 52.10%. This paper revealed erization selectivity of linear alkanes, and gave their interpretation.
the relation between the Brønsted acid strength and the Al atom coor­ With the combination of theoretical calculation (DFT) and experiments,
dination properties, and confirmed that the mild acid strength favored n-heptane and its isomers were chosen as model reactants, and alumi­
the high selectivity. Parmar et al. [61] reported hydroisomerization of n- nosilicates with various pore topology and acid strength (tuned by
hexadecane over Pt/ZSM-22 catalysts, in which the acidity of ZSM-22 different heteroatom substitution in the framework) were chosen as acid
was tuned by ion-exchange method. Using Group II cations, Ca2+, supports. From the results, it was demonstrated that the intrinsic
Mg2+, Ba2+, in the ion-exchange of ZSM-22, remarkable drop in β-scission/isomerization rate ratio on each single-sojourn acid site
Brønsted acid sites and reduction of Brønsted acidity level can be decrease with the increase of acid strength. With the decrease of acid
observed. The hydroisomerization of n-hexadecane over Pt/ZSM-22 strength, β-scission transition states became more preferentially stabi­
catalysts using calcium-exchanged ZSM-22 zeolites showed evidently lized. Contradict to general impact β-scission was favored on mild or
higher selectivity of branched isomers. For example, Ca2+ exchanged weaker acid sites rather than isomerization. It was argued the observa­
catalysts showed Brønsted acid sites decreased by 20.4%, and the iso­ tion that hydroisomerization of long-chain alkanes on catalysts with
mers selectivity increased by 6–7%. The ion-exchanged ZSM-22 was said strong acid strength led to higher cracking was due to the diffusion-
to favor better metal and acid balance in catalysts. The mild acid enhanced secondary β-scission of primary isomers. Therefore, whether
strength of catalysts after ion exchange favored the isomerization rather to enhance or reduce the acid strength for higher yields of isomers still
than cracking of adsorbed carbenium ions [62]. Feng et al. [63] studied needs to be discussed. And in the future studies, it shall be accompanied
the effects of dynamic and static crystallization in the synthesis of ZSM- with insights to the diffusion constraints.
22 zeolites. Using different crystallization conditions, ZSM-22 zeolites From above researches, it can be concluded that the tuning of acid
were formed from precursors with the same SiO2/Al2O3 ratios from 70 strength and density in hydroisomerization of light alkanes and long-
to 110. The analysis of these ZSM-22 samples showed that the differ­ chain alkanes have great difference. For light alkanes, the goal of
ences are mainly exhibited on two aspects, namely Brønsted acidity and enhancing isomer yields was always related to the increase of acid
crystal size. At each SiO2/Al2O3 ratio, those ZSM-22 samples crystallized strength and the strong acid density. It seems the isomerization reaction
in the dynamic condition showed lower overall Brønsted acid sites for n-pentane and n-hexane was favored on Brønsted acid sites with
density, especially for those strong ones. The Brønsted acidity difference higher strength. However, for long-chain alkanes, the modifications
was attributed to the higher framework Si/Al ratio and the decrease of were aiming at reducing the acid strength and density. This significant
Si-O-Al bridge hydroxyl. The hydroisomerization of n-Decane using difference mainly addresses from two reasons. The first reason is the
catalysts based on dynamically formed ZSM-22 showed higher isomer reactivity of long-chain alkanes is higher, therefore, mild acid strength is
selectivity. It was attributed to the relatively weak Brønsted acidity and enough to catalyze the reaction. This discrepancy in reactivity is mainly
more pore mouth exposure. manifested in the overall conversion versus temperature under the same
Niu et al. [64] reported the synthesis of siliceous ZSM-22 zeolite, condition. For example, as shown in Fig. 1, to reach the same conversion
named Z-Si. The analysis of Z-Si and comparison with silica–alumina level, the temperature needed for n-hexane, n-heptane and n-octane
ZSM-22 showed significant reduction of Brønsted acid density and follows the sequence of n-hexane > n-heptane > n-octane [67]. Similar
strength. The total Brønsted acid density was reduced from 176 μmol g− 1 results were also reported from n-C5 to n-C11 [57,68–70]. The second
on Z-40 to 13 μmol g− 1 on Z-Si, and the density of strong ones was reason is alkanes with longer chain length suffer more severe diffusion
reduced from 146 μmol g− 1 to 3 μmol g− 1. The lowered acidity was limitations [39]. In Fig. 2, Smit [39] has summarized the self-
attributed to the existence of vicinal silanol and silanol nests in Z-Si. Pt- coefficients of n-alkanes of different chain length according to the re­
loaded bifunctional catalysts were prepared with Z-Si and silica­ views of Jobic and Theodorou [71,72]. It can be seen whether it is from
–alumina ZSM-22. Hydroisomerization of n-Dodecane using these cat­ molecular simulation or experiment, the diffusion coefficient decreases
alysts showed the selectivity was higher by about 10% on Pt/Z-Si almost exponentially with the increase of chain length. If the acid
catalyst at each conversion level. These results can be attributed to the strength is too strong, excessive isomerization would happen during the
low Brønsted acid density and weak strength introduced by the vicinal slow diffusion. As a result, cracking reactions will be dominant. How­
silanol and silanol nests. The weak acid strength favors the desorption of ever, the lower acid strength also causes excessive energy consumption.
isomers and can suppress the successive isomerization, therefore limit For example, to reach a conversion of 50%, a reaction temperature of

6
C. Wei et al. Fuel xxx (xxxx) xxx

Fig. 1. Conversion of individual alkanes (nC6, nC7 and nC8) vs. the reaction temperature on a Pt (a) or Pd (b) loaded H-Beta catalyst. Adapted from [67] with
permission from Elsevier.

catalysts, efforts have been made to tune the location of acid sites.
To date, ZSM-22 is one of the most widely used and studied acid
supports in the hydroisomerization of long-chain alkanes. It’s argued
that the Brønsted acid site locations for ZSM-22 zeolites can be generally
classified as three types: on the external lateral surface of ZSM-22 rods,
on the external surface close to the pore mouth and inside the micro­
pores [77]. Those acid sites inside the micropores is claimed only
accessible for n-alkanes because of the narrow pore size of ZSM-22,
which is 0.45*0.55 nm [78]. Furthermore, n-alkanes which is going
through the micropores usually undergo β-scission due to the severe
spatial confinement. While the acid sites on the lateral surface are
usually regarded have no shape selectivity, so they will randomly absorb
branched isomers and thus cause β-scission. Only those acid sites near
the pore mouth are believed have shape selectivity due to “pore mouth”
and “key lock” mechanisms as mentioned above.
Wang et al. [79] synthesized a series of Pt/ZSM-22 catalysts with
different acid densities on different locations. In their work, two stra­
tegies were used to tune the acidity. The first strategy was based on the
protective effects of the templates during NH4+ ion exchange.
Compared to conventional fully-detemplated ZSM-22, the micropore
volume of partially-detemplated ones was reduced from 0.090 to 0.094
Fig. 2. Self-diffusion coefficients of the n-alkanes in MFI as a function of chain
cm3/g to 0.001–0.007 cm3/g. And the retained template was believed to
length at 300 K as obtained by various techniques: ●, molecular simulations; □,
have exhibited the NH4+ ion exchange within micropores, and thus
QENS; △, PFG-NMR. The asterisks correspond to extrapolations to 300 K.
Adapted from [39] with permission from ACS.
reduced the acid density inside the micropores. The second strategy was
tailoring the total Brønsted acid density by adjusting the concentration
of NH4+ in the ion exchange step. It’s shown that with the decrease of
200–250 ◦ C is required for n-pentane and n-hexane on MOR based
NH4+ concentration, the total acid density was also reduced, without
catalysts, a higher temperature of 300–400 ◦ C is necessary for long-
changing the relative acid distribution. The change of total acid density
chain alkanes on ZSM-22 and SAPO-11 based catalysts [23,58,73,74].
and accessibility factor (AF) caused by these strategies was shown in
Not to mention the hydroisomerization is equilibrium limited, which
Fig. 3. The accessibility factor (AF) was defined as the ratio between the
means high yield for isomers at higher temperature is not allowed
amount of external Brönsted acidity and total Brönsted acidity [80]. In
[68,75,76]. Thus, the mild acid strength for hydroisomerization of long-
the end, they made the catalyst Pt/230AN0.2 by combining these two
chain alkanes can be seen as a sacrifice to the diffusion limitations.
strategies. The catalytic hydroisomerization of n-dodecane on Pt/
230AN0.2 showed this combined strategy achieved the highest isomer
3.2.2. Effects of acid sites location
yields increased from 40.7% to 63.7%. It is regraded that there exists a
To the best of our knowledge, research about the effects of acid sites
kind of synergic effect with the two strategies. The reduced acid density
location for light alkanes is rare. However, its importance for long-chain
inside the micropores exhibited cracking reactions of n-alkanes diffusing
alkanes is already confirmed by some researchers. Therefore, this sub­
along the channels. While the reduced total acid density not only limited
section is mainly focused on long-chain alkanes. For long-chain alkanes,
the cracking on the external surface, but also increased the isomeriza­
it is believed only linear ones can enter the micropores rather than
tion through “key lock” mechanism.
branched ones. Indeed, it is one kind of shape selectivity and is broadly
Niu et al. [81] studied the effects of acid site distribution of ZSM-22
utilized to enhance isomer yields. However, because of the severe
on the isomers selectivity. Firstly, they synthesized a ZSM-22 with a
diffusion constraints imposed by the micropore channels, excessive
core–shell structure to tune the acid function on the external surface. In
contact time between isomers and Brønsted acid sites tends to cause
their research, the epitaxial growth method was used to fabricate the
secondary cracking. Besides, nonselective external acid sites also favor
highly siliceous shell with reduced acid density on the external surface,
secondary cracking. To deal with the secondary cracking of isomers
namely reduced by more than 50% compared to conventional ones.
trapped in the micropores or adsorbed on the external surface of

7
C. Wei et al. Fuel xxx (xxxx) xxx

Fig. 3. Correlation between Micropore Exposed and total Brönsted acidity & AF (a), and correlation between NH4+ aqueous concentration and total Brönsted acidity
& AF (b), and schematic acidity distribution. Adapted from [79] with permission from Elsevier.

Unexpectedly, using catalysts based on this ZSM-22, isomers selectivity the acid sites of SAPO-11 were predominantly located on the external
in n-dodecane hydroisomerization was found reduced by about surface. While the use of carbon templating maintained the predominant
0.9–6.4% at the conversion near 90%, compared with catalysts based on acid location inside the micropores, and fabricated mesopores structure.
conventional ZSM-22. It was attributed to the reduced external acid The result showed that the maximum isomer yields of n-dodecane on the
density suppressed the adsorption of reactants on the lateral surface, catalyst with mainly external acid sites was only 13%, while that of the
thus in turn enhanced the reactants’ diffusion inside the micropore. And catalyst with predominantly internal acid sites was 84%. In comparison,
the diffusion confinement, as well as the acid catalysis inside the pores, this number in conventional Pt/SAPO-11 without mesopores structure
caused excessive cracking. Therefore, another strategy was taken to was 44%. The results indicated that the nonselective external acid sites
study the effects of acid sites inside the micropores. They used carbon­ on SAPO-11 mainly favors the cracking reaction of isomers, while the
ated templates to block the micropore of core–shell ZSM-22, as shown in internal acid sites are selective for isomerization. However, the effect of
Fig. 4. Using n-dodecane as model reactant, the hydroisomerization removing diffusion constraints was also obvious for higher isomer
result showed that the isomers selectivity was enhanced by 2.2–2.7% yields. Zhang et al. [83] reported the preparation of a kind of stable
compared with catalysts based on conventional ZSM-22, at the conver­ SAPO-11 nanosheets with partially blocked acidic sites inside micro­
sion near 90%. Since the Brönsted acid density on the external surface pores, named as N-SAPO-11. The synthesized N-SAPO-11 had the
was reduced to a relatively low value, and acid sites deep inside the thickness of 10 to 20 nm, and the acidic sites inside micropores were
micropores were not accessible due to the blockage, the acid sites near partially blocked. Therefore, the acid sites near the pore mouth became
the pore mouth thus became dominant in the acid function. Similar predominant. Using the Pt loaded N-SAPO-11 and conventional bulk
result was also obtained by Lv et al. [82]. On Pt/Z-22PF catalyst with SAPO-11 as catalysts, the reseult of n-dodecane hydroisomerization
ZSM-22 was partially filled channels by in-situ coke deposition, the showed evident difference on isomers selectivity. Catalyst based on N-
maximum isomer yield was enhanced by 6% compared to conventional SAPO-11 always had higher isomers selectivity by about 5% at each
Pt/ZSM-22 catalyst. The increased selectivity was attributed to the conversion. The enhanced isomers selectivity was attributed to the in­
highly selective “pore mouth” and “key lock” catalysis. crease of isomerization through “pore mouth” and “key lock” catalysis.
The same acid location effects can be observed on SAPO-11 molec­ In this section, the effects of acid sites location were discussed. First,
ular sieves. Kim et al. [80] reported the hydroisomerization of n- the key role of acid sites location on catalysts for hydroisomerization of
dodecane on mesoporous Pt/SAPO-11 catalysts with different predom­ long-chain alkanes was proved. Acid sites located on the external surface
inant locations of acid sites. With the help of organosilane templating, in some degree are responsible for secondary cracking, due to the lack of

Fig. 4. A schematic diagram showing the structure of core–shell ZSM-22 with micropores blockage. Adapted from [81] with permission from RSC.

8
C. Wei et al. Fuel xxx (xxxx) xxx

shape selectivity. While the acid sites near pore mouth are in favor of pore Y. Similar work and results were reported by Ben Moussa et al.
isomerization through “pore mouth” and “key lock” catalysis, which are [86], using Pt loaded aggregates of 12-ring pore BEA zeolite and
highly selective. Therefore, high isomer yields can be obtained when the alumina. In their research, nanoscale and microscale aggregates were
acid sites near pore mouth become predominant. However, the effects of prepared, and Pt was selectively loaded on zeolite or alumina to give a
acid sites inside micropores are not in good agreement. Some studies closer or farther intimacy between metal sites and acid sites. Obviously,
tend to believe its negative influence on isomers selectivity, considering no matter the size of aggregates, when Pt was loaded on zeolites, the
the diffusion constraints inside micropores may prolong the contact time closer metal–acid intimacy was obtained. However, hydroisomerization
between primary isomers and acid sites, and thus increase secondary results of n-heptane showed catalyst with Pt located on alumina
cracking. While others tend to regard it is benefit for selectivity possessed higher isomerization selectivity at conversion at 20.5 ± 2%
enhancement. In our view, the effects of acid sites located inside mi­ and 50 ± 4%, rather than the one with Pt located on zeolites. In other
cropores may vary based on the structure of acid supports and the chain words, the best hydroisomerization performance was achieved on the
length of reactants. Therefore, further research is necessary to acquire catalyst with farther intimacy other than on the one with the closest
valid conclusion. intimacy in the range of nanometer scale. These unique behaviors were
further explained by Oenema et al. [87] using so-called “remember ef­
3.3. Effect of metal–acid intimacy fect” that when the metal sites were not located inside zeolites micro­
pores, e.g. on alumina binder, linear alkene intermediates will be
Considering the hydroisomerization reaction is related to reactant catalyzed in the outer layer of zeolites and remember the entrance to
diffusion between hydrogenation function and acid function, it is timely escape to avoid further reaction. Of note is that, in above-
naturally to question how is the reaction affected by the distance of these mentioned studies, it has been emphasized the metal–acid intimacy
two functions. closer than nanoscale was obtained by placing Pt particles inside
Weisz et al. [84] have proposed the so-called “intimacy criterion” for channels of zeolites. In such condition, the metal–acid intimacy is not
hydroisomerization by using physically mixed Pt/silica and acidic silica- the only variable anymore, the change of pore structure of zeolites has
alumina catalysts with different size in the 1962. They proved that in the also been included implicitly.
range of micrometer scale and millimeter scale, isomer yield of n-al­ To exclude impacts imposed by pore structure, nonporous materials
kanes hydroisomerization increase with the decline of metal–acid sites were used to study the effect of metal–acid site proximity by Samad et al.
distance. Beyond millimeter scale distance, the bifunctionality of the [88] Four catalysts with different metal–acid intimacy level were pre­
catalysts almost disappeared. This phenomenon was ascribed to the pared, namely PTA-11 and PTA-DI with atomic-scale, CPA-4 with
limited diffusion length of alkene intermediates beyond which they have nanometer-scale, PM with micrometer-scale, and L-1 with millimeter-
little chance to access another metal site to be hydrogenated. It was scale. The atomic-scale and nanometer-scale intimacy were achieved
further simplified to the principle “the close the better” for metal–acid by selectively adsorption of Pt precursors on silica-alumina mixed oxides
intimacy. Their work was recently re-confirmed by Batalha et al. [51] or on silica. Micrometer-scale and millimeter-scale affinities were con­
They studied the effect of metal–acid intimacy in the range of micro­ ducted by physical mixing or quartz separation of Pt loaded silica and
meter distance using S1, S2 and S3 series catalysts which are based on acidic silica-alumina. The results of n-heptane hydroisomerization
zeolites and alumina. The intimacy difference was possessed by BEA showed the catalyst CPA-4 with nanometer-scale possessed ~13%
zeolites ion exchanging with Pt precursor or mixing with Pt loaded higher i-C7 selectivity than atomic-scale catalysts at conversion 22 ± 3%
alumina. The distances between acid site to metal site on S1, S2, and S3 conversion. This proves that even if the change in pore structure is
were estimated by the size of aggregates with at least 12.5 μm, 12.5 excluded, at nanometer and smaller scales, principle of “the closer the
μm–300 μm and 300 μm, respectively. It was found the yield of i-C16 in better” is still not applicable.
n-C16 hydroisomerization decrease with the sequence of S1 > S2 > S3 However, the opposite results were obtained by Mendes et al. [89] in
with ideal CPt/CH+ ratio. Besides, turn-over frequency (TOF) of the acid 2020. They also prepared composite catalysts with zeolite Y and
sites varies in the same trend. These results indicate that in the intimacy alumina, selectively deposing Pt on one of them to acquire different
range they studied, the catalytic performance still follows the principle distance between metal and acid sites. On Pt/HUSY-Al2O3 with
of “the closer the better”. However, it has been being challenged in the maximum acid-Pt distance of 10–20 nm, the maximum yield of i-C16 in
last few years. n-C16 hydroisomerization was much higher than it on Pt/Al2O3-HUSY
In 2015, Zečević et al. [85] studied the metal–acid intimacy in with acid-Pt distance of 100–10000 nm. Their work supported “the
nanometer scale. The catalysts they used were Pt loaded extrudes of closer the better” principle again. And they have pointed out the loop­
γ-alumina binder (negligible acidity) and Y zeolites (acid site supplier). holes of the explanation given by Zečević et al. [85] for Pt loaded
The Pt was selectively loaded on γ-alumina or on Y zeolites through bifunctional catalysts with Pt on alumina binder, i.e. larger metal–acid
choosing Pt precursors with different electric properties and adjusting distance, possessing higher isomers yield. They claimed if the place
pH of solutions, denoted as Pt-A/Y and Pt-Y/A respectively. HAADF- where the isomerization reaction occurs is zeolites’ outer layer, then the
STEM and EDX were used to characterize catalyst slices with 70-nm- conversion of the reaction should be reduced. However, that was not the
thick, and the closest metal–acid intimacy of Pt-Y/A and nanometer- fact.
scale intimacy of Pt-A/Y was confirmed. Hydroisomerization results of In conclusion, the metal–acid intimacy of bifunctional catalysts plays
n-C10 and n-C19 both showed higher isomerization selectivity on Pt-A/ a crucial role in hydroisomerization of n-alkanes. Once the distance
Y with identical conversion, being contrary to “the closer the better” between metal and acid sites is large enough, the activity will be reduced
principle. It was attributed to that Pt particles were mainly located in­ dramatically for the two functions cannot cooperate or the isomer yield
side the micropores to reach the closest intimacy in the case of Pt-Y/A, will decline for undesired secondary reactions. However, no unified
which means alkene intermediates have to go through the zeolite point of view has been drawn on how far the intimacy should be and it
channels to access another metal site. In this process, it will suffer from needs further study.
diffusion limitations and lead to higher possibility of secondary re­
actions of cracking and coking. On the other hand, on Pt-A/Y catalyst, 3.4. Recent advance in modifying metal sites
alkene intermediates could be isomerized on the surface of zeolite Y and
diffuse back to the metal sites on alumina. Their work not only supplied Noble metals like Pt and Pd are the most popularly used components
counterexample to the conventional intimacy criterion, but also to endue hydroisomerization catalysts (de)hydrogenation ability in
extended the view of the long chain alkene intermediates were mainly theoretical research and industrial application. However, their high
isomerized in the outer layer of zeolites from medium-pore ones to large- price has also caused worry in expanding their use. Recently, efforts

9
C. Wei et al. Fuel xxx (xxxx) xxx

have been made on lowering their content, enhancing their hydroge­ performance can be obtained on the catalyst with 13 wt% Mo loading
nation efficiency and replacing them with non-noble metals, accompa­ through physical mixing of MoO3 with ASA. The i-C7 yield on this
nied by the introduction of new loading methods. catalyst was comparable to the counterpart of 0.7 wt% Pt loading
It’s generally believed high hydrogenation efficiency of Pt and Pd (64.5% versus 65.7%). It was believed that MoO2 with a crystal size of
can be reached by small particle size and high dispersion for the unique 30–40 nm was the active phase supplying hydrogenation function, and is
electron structure of surface metal atoms and superior activity of some coordinated to the surface oxygen and MoOx nearby. However, consid­
crystal facets, thereby, their minimum content in catalysts can be ering the extremely high loading of Mo, the application of this meth­
reduced in some degree [90–92]. In this regard, Wang et al. [93] odology in the most commonly used zeolite-type catalysts might
developed the “vacuum assistant” strategy (VA) to enhance the disper­ encounter problems, such as severe blockage of micropores. To avoid the
sion of Pt clusters on SAPO-11, thereby lowering the loading. The key blockage addressed by overloading and at the same time enhance the
step was to exhaust gas from the micropores of SAPO-11 before dispersion of non-noble metal, Kim et al. [43] reported replacing Pt with
impregnating the metal precursor. In this way, they made V-0.15Pt/S-11 2 wt% Ni in hydroisomerization. In their research, a key catalysis ma­
catalyst, namely SAPO-11 loaded with 0.15 wt% Pt. The counterpart, terial, namely nanosponge-like ZSM-5, was synthesized with the help of
catalyst C-0.5Pt/S-11 was made through conventional impregnation a multiammonium surfactant as a structure-directing agent. Its unique
procedure with Pt loading of 0.5 wt%. The hydroisomerization perfor­ morphology was constructed by ZSM-5 nanosheets with 2.5 nm thick­
mance of n-C16 has shown that V-0.15Pt/S-11 was superior to C-0.5Pt/ ness along the 10-ring straight channel, and mesopores of ca. 5 nm were
S-11, with a maximum i-C16 yield of 89% vs. 80%. The author ascribed evenly distributed between these sheets. Nickle was loaded on this ZSM-
the superiority to its smaller Pt particle size (~1 nm) and better synergy 5 by incipient wetness impregnation, and 1 wt% Pt was loaded to be
effect with acid sites achieved by the VA strategy. Intensive capillary compared. In the hydroisomerization of n-dodecane, the Ni-loaded
action of Pt precursor in the “clean” channels of SAPO-11 guaranteed catalyst with optimized content (2 wt%) showed comparable activity
their effective adsorption, which was hindered by gas resistance inside and isomer yield to the Pt-loaded counterpart. The high efficiency of Ni-
through conventional methods. Geerts et al. [94] applied the atomic loaded catalyst came from its small particle size of ca. 5.4 nm. This high
layer deposition (ALD) method to deposit Pt nanoparticles on ZSM-5. dispersion of Ni on conventional ZSM-5 can hardly be achieved since the
Trimethyl(methylcyclopentadienyl)platinum was chosen as Pt precur­ relatively weak interaction between Ni species and the support will
sor in cold wall ALD. Compared to conventional incipient wetness cause severe agglomeration during calcination and H2 reduction.
impregnation, a smaller size of Pt particles was obtained (2–4 nm versus However, thanks to the uniform mesopores in nanosponge-like ZSM-5,
15–20 nm). Therefore, the reactivity and isomer selectivity of the the Ni species can be stabilized by mesopore walls like a sandwich.
catalyst was enhanced in n-decane hydroisomerization. Al-Rawi et al. Therefore, the strengthened interaction prevented agglomeration during
[95] reported a supercritical CO2 (ScCO2) assisted impregnation the reduction process. This idea of controlling the structure of the sup­
method, using dimethyl(1,5-cyclooctadiene) platinum as the precursor. port to enhance the its interaction with metal species, thereby improving
For comparison, the conventional wetness impregnation method was the hydrogenation efficiency of non-noble metals is promising.
used. Based on H-ZSM and H-MOR, they assessed the influence of the Pt From the above discussion, we can see that people have made efforts
loading method on n-hexane hydroisomerization. It was found the cat­ to reduce the amount of precious metals from different aspects,
alysts loaded by ScCO2 showed higher isomer yield and lower cracking including adopting new loading methods, regulating non-precious
yield at all temperatures. Besides, for catalysts loaded by ScCO2, the metals to completely replace precious metals, and so on. These efforts
cracking reaction was much less sensitive to temperature change are of great significance to the industrial promotion of short-chain and
compared to wet impregnated ones. The superiorities were ascribed to long-chain n-alkane hydroisomerization technology and the production
the smaller and more uniformly distributed Pt particles by ScCO2 (1–2 of low-cost clean oil products.
nm vs. 6–9 nm). The function of ScCO2 in the loading process was
believed to have accelerated the nucleation rate and suppressed the 4. Modifications on the textural structure of molecular sieves for
growth rate of Pt crystals, and finally the Pt size and distribution were long-chain alkanes
affected.
In addition to enhancing the hydrogenation efficiency of noble metal Molecular sieves-based catalysts for hydroisomerization show great
sites by adjusting loading methods, attempts have also been done to shape selectivity for isomers. Since for long-chain reactants, the modi­
entirely replace noble metal by non-noble metals. Wang et al. [96] have fications on the textural structure of molecular sieves are much more
explored the application of non-noble bimetallic catalysts in hydro­ important than it for light alkanes, due to their worse diffusivity and
isomerization. The same “vacuum assistant” strategy was used to co- higher reactivity. Meanwhile, modification of textural structure for light
impregnate nickel and molybdenum on SAPO-11. A series of catalysts alkanes isomerization is rarely reported. The effects of the textural
based on SAPO-11 with different Ni and Mo loadings were made to structure modifications only for the hydroisomerization of long-chain
optimize the metal content. The conventional 0.5 wt% Pt loaded SAPO- alkanes will be discussed in this chapter. The most used acid supports
11 was chosen to be compared. In the hydroisomerization of n-C16, are unidimensional 10 MR or 12 MR microporous materials, such as,
3.0Ni-0.5Mo/S-11 was found possessing comparable maximum i-C16 ZSM-22 and SAPO-11, considering their unidimensional channels have
yield with Pt loaded catalyst, namely 81.4% versus 81.0%. Through great reactant selectivity for linear alkanes. The diffusion constraints
their analysis, several origins were found to have gifted 3.0Ni-0.5Mo/S- inside the channels are always severe due to the unsuitable path length,
11 high performance. Firstly, the addition of Mo enhanced the interac­ and it leads to undesired cracking. To overcome this disadvantage and
tion between Ni and SAPO-11, resulting in its smaller particle size and increase the pore mouth exposure, efforts have been made mainly in two
higher dispersion during reduction, which is positive for enhancing aspects, i.e., fabricate hierarchical pore structure and reduce the crystal
hydrogenation ability [97]. Besides, the existence of MoOx species has size of catalytic materials.
drawn electron transfer between Mo and Ni due to orbital overlapping
and the tuning of surface charge was believed to enhance the intrinsic 4.1. Effect of hierarchical pore structure
activity of Ni, resulting in easier H activation and desorption. Harmel
et al. [98] showed the possibility of replacing Pt with molybdenum Hierarchical pore structure refers to the fabricated mesoporous and/
oxide. They prepared a series of catalysts with MoOx species acting as or macroporous channels in the microporous structure. Compared to
hydrogenation function, and the acid function was supplied by amor­ pure micropores, hierarchical pore structure is regarded to have better
phous silica-alumina (ASA) or Al-SBA-15. By adjusting the loading diffusivity and shortened diffusion length. There are mainly two series of
method and content, they found the best n-C7 hydroisomerization methods used to fabricate the hierarchical pore structure, which are so

10
C. Wei et al. Fuel xxx (xxxx) xxx

called “top down” method and “bottom up” method. Besides the conventional methods, researchers have made great ef­
“Bottom up” method means by using some physiochemical forts on controlling the crystallization step of synthesis gel or precursors
methods, the hierarchical structure is fabricated directly in the synthe­ to attain hierarchical pore structure. Zhang et al. [103] reported a two-
sis. Using mesoporogen and template to fabricate hierarchical structure stage hydrothermal crystallization route to synthesize mesoporous
can be regarded as a conventional “Bottom up” method. However, the SAPO-11 with two types of mesopores, smaller ones and bigger ones,
most disappointed feature of “Bottom up” method is that it cannot be centered at ca. 2.4 nm and at ca. 17–24 nm. The fabrication of the micro-
applied to all the catalytic materials. Nevertheless, over last few years, mesoporous structure was achieved by the addition of CTAB and F127.
studies on suitable hierarchical pore structure catalysts of hydro­ The result of n-C8 hydroisomerization showed maximum isomer yields
isomerization based on this method are still attractive. promotion from 60.9% to 73.4%. The promotion was attributed to the
Fan et al. [99] reported the hydrothermal synthesis of intersecting shortened path length and increased pore mouth, which favors highly
micropore and mesopore structural SAPO-11 by simultaneously using selective “pore mouth” and “key lock” catalysis. Tao et al. [104]
tetradecylphosphoric acid (TDPA) as mesoporous template and di-n- demonstrated a confined space approach to fabricate hierarchical
propylamine (DPA) as microporous template. The characterization of MgAPO-11 with AEL type structure which is the same as SAPO-11. By
high-resolution transmission electron microscopy demonstrated the adding furfuryl alcohol (FA) into mesoporous magnesioaluminophos­
formation of mesopores with 4–7 nm pore size. N2 adsorp­ phate (MAP) and successively carbonating FA into carbon film, the
tion–desorption showed the mesopore volume of hierarchical SAPO-11 coated carbon film can supply the space confinement during the crys­
increased from 0.06 cm3/g to 0.31 cm3/g compared with conventional tallization of MAP. As a result, the hierarchical MgAPO-11 with meso­
sample. The catalyst was prepared by loading Pt on the hierarchical pores centered at 7 nm and 20 nm was formed. Comparison of n-
SAPO-11 sample. The result of n-C8 hydroisomerization on this catalyst dodecane hydroisomerization was made between catalysts based on
showed the total selectivity of iso-C8 was increased from 88.5% to 97.9% hierarchical MAP and microporous MAP. The result showed higher
at the conversion of 43%, and the selectivity of di-branched iso-C8 was isomer yield was obtained at 90% conversion on hierarchical one
increased by 21.5%. To explain the increased isomers selectivity, espe­ (86.6% vs 60.5%). The promoted isomer yield was attributed to the
cially for di-branched ones, they proposed a novel reaction path, which shortened path length enhanced diffusion of multibranched isomers out
is shown in Fig. 5. It claimed that the mono-branched isomers were of micropores.
mainly formed inside the micropores, however, the fabricated hierar­ Recently, a novel method of affecting the crystallization and intro­
chical structure shortened their path length, thus reduced the possibility ducing mesopores in molecular sieves was applied, which is called the
of cracking. Successively, due to the mutually penetrating between mi­ oriented assembly strategy. Jin et al. [105] reported the synthesis of
cropores and mesopores, those mono-branched isomers were further hierarchical SAPO-11 by the oriented assembly of prefabricated nano­
isomerized into di-branched form without steric hindrance. In the end, crystallites. The SAPO-11 building blocks were prefabricated and fol­
the isomers selectivity was promoted. This newly proposed reaction path lowed by further crystallization by particle attachment. The record
expanded the roles of mesopores in the isomer selectivity enhancement. hydroisomerization performance was obtained on SAPO-11-PDDA,
Tao et al. [100] reported the synthesis of inter-crystalline meso­ whose hierarchical structure was confirmed with mesopores volume
porous SAPO-11 using dry gel conversion, [3-(Trimethoxysilyl)propyl] increased from 0.13 cm3 g− 1 to 0.26 cm3 g− 1, BET surface area increased
octadecyldimethylammoniumchloride (TPOAC) was added as meso­ from 134 m2 g− 1 to 302 m2 g− 1 compared with conventional SAPO-11.
pores template. In as-synthesized SAPO-11, mesopores centered at 4.5 The result of n-C7 hydroisomerization on Pt/SAPO-11-PDDA catalysts
nm were fabricated. Catalysts were made by loading Pt. The result of n- showed a record yields (79%) of isomer. The advanced isomers selec­
dodecane hydroisomerization showed that the isomer yields increased tivity was attributed that the enhanced diffusivity and widened intrinsic
from 69.2% to 77.3% at the conversion of about 90% compared with on reaction-controlled regime successfully reduced the cracking possibility.
catalysts based on conventional SAPO-11. Recently, they have devel­ In 2018, they [106] reported another study of mesostructured SAPO-11
oped a new way to synthesize hierarchical nanosized single-crystal synthesis by oriented assembly strategy. The same method was used for
SAPO-11 without using mesoporogen [101]. By carefully adjusting the preparing the SAPO-11 building blocks [105]. In the assembly process,
composition of precursors and heating at elevated temperatures, the they demonstrated the formation of “house of cards” structure with
obtained SAPO-11 was demonstrated to have ~65% of intracrystalline mesopores that can be attained under the dynamic tumbling condition
mesopores accessible to the crystal surface, which is promising to without any additives. Besides, the addition of growth modifier cellulose
enhance hydroisomerization performance. Wen et al. [102] reported the 2-(2-hydroxy-3-(trimethylammonium) propoxy) ethyl ether chloride
synthesis of mesoporous Pt/SAPO-11 catalysts using F127 and super­ (polyquaternium-10 or PQ-10) in the assembly process can greatly
critical carbon dioxide as combined mesoporous templates. The resulted reduce the size of SAPO-11 nanosheets from ca. 300 nm to ca. 20 nm.
hydroisomerization reaction of n-C9 showed the high selectivity of di- BET surface area and mesopores volume of tumbling assisted (SAPO-11-
branched isomers of 33.5% and low cracking selectivity of 8.0% at the T) and PQ-10 assisted (SAPO-11-PQ) samples were increased by 41.0%-
conversion of 20%. The enhanced selectivity in the above two pieces of 78.4% and 23.1%-123.1% compared to conventional one (SAPO-11-C),
research was attributed to the reduced diffusion constraints by short­ respectively. Fick’s second low was used to quantitatively measure and
ened diffusion path. And the use of mesopores templates can be regarded compare the diffusion properties of SAPO-11-C, SAPO-11-T and SAPO-
as an effective route to synthesize mesoporous molecular sieves. PQ. The diffusion time constants of 2-methylhexane on SAPO-11-C,
SAPO-11-T and SAPO-PQ were 0.77 × 10− 3, 0.99 × 10− 3, and 1.17 ×
10− 3, respectively. As can be seen, SAPO-11-PQ has the best diffusion
properties, consistent with the highest iso-C7 yield of ca. 80% on Pt/
SAPO-11-PQ. These studies firstly confirmed the positive effects of hi­
erarchical pore structure on hydroisomerization of long-chain alkanes,
and indicated the novel oriented assembly strategy is not only a prom­
ising but also a flexible route to synthesis hierarchical pore structure as
well.
These results suggest that due to the hierarchical pore system con­
structed by both micropores and mesopores, the enhancement of isomer
Fig. 5. Reaction and diffusion pathways for n-octane isomerization in the pore selectivity for long-chain alkanes is remarkable. However, if one wants
channels of H-SAPO-11-HI monocrystal. Adapted from [99] with permission to fabricate molecular sieves with hierarchical structure by “bottom up”
from Elsevier. methods, the efforts on controlling the crystallization step should be

11
C. Wei et al. Fuel xxx (xxxx) xxx

paid. Furthermore, the usage of mesoporogen seems to be necessary at


most conditions. Therefore, development of novel routes in which the
consumption of mesoporogen can be reduced or avoided is favored in
the future.
“Top down” method means using the fresh molecular sieves as the
parent materials, and remodeling it by post-treatment strategies. In the
past few decades, people have tried dealumination with acid solutions to
post-treat molecular sieves to attain mesopores [107,108]. However,
using acid solutions will also cause undesired metal leaching and thus
influence acid function [109]. Therefore, over the last few years, re­
searchers paid more attention on alkaline desilication. The general
alkaline desilication procedures can be described as follows: First, pre­
pare alkaline water solutions with proper alkaline species and concen­ Fig. 6. Brief mechanism scheme of ZSM-22 dual-protected alkali treatment.
trations. In common, people use NaOH as alkaline species, and different Adapted from [117] with permission from ACS.
additives are added into the solutions, such as ethanol and surfactants.
Second, disperse the fresh molecular sieves into the alkaline solutions, the Si leaching process by drilling effect and result in uniform meso­
and maintain hydrothermal reaction at middle temperature for hours. In porosity whose size is close to the CTAB micelles. The result of n-C12
the end, these treated molecular sieves should be washed with water or hydroisomerization showed whether ZSM-22 was desilicated with or
dilute acid solutions. without protection, total isomers yield on the mesoporous catalyst was
The most important influencing factors of alkali desilication were increase by 4.8%–33.6%. Besides, higher mono-branched and total
studied by Wei et al. [110] Their research was based on ZSM-12 zeolite, isomers yield was obtained on dual-protected ZSM-22 samples. The
and factors of NaOH concentration, alkaline treatment temperature, and enhanced catalytic performance on desilicated ZSM-48 and ZSM-22
period were studied. The NaOH concentration was regarded as the most samples was attributed to the increased pore mouth exposure and bet­
influential in the desilication. Below the NaOH concentration of 0.2 M, ter diffusion properties. While the better performance on samples with
no mesopore fabrication can be observed. Higher temperature is favored protected-desilication process was addressed to that the protection
in desilication; however, its effect was not comparable to the effect of agents have overcome the uncontrollable silicon extraction.
concentration. While beyond 90 min period of desilication, changes in Liu et al. [118] reported the synthesis of hierarchical ZSM-22 acid
the desilication rate can hardly be found anymore. The study suggested support using the alkaline treatment and recrystallization process. In
that there exists a best NaOH concentration range and a proper treat­ their experiments, the H-ZSM-22 molecular sieves were treated with
ment period in generating mesopores in molecular sieves. And in further different concentrations of alkaline solutions and recrystallization with
studies of desilication of molecular sieves, all these factors should be the help of CTAB. In comparison, they also synthesized the desilicated
considered to proceed the experiments. Beyond the desilication condi­ samples using alkaline treatment alone, which was denoted as DeZEO.
tions, they also studied the desilication process of ZSM-12 with different The study of the textual properties showed that the samples fabricated
Si/Al ratios. It’s found that, in the same desilication condition, those from simultaneous alkaline treatment and recrystallization process
molecular sieves with higher Si/Al ratios favored the fabrication of attained uniform mesoporous materials on the lateral surface, while
mesopores. It was attributed to that the framework Al can hinder the DeZEO sample attained uniform mesopores. The uniform mesoporous
extraction of Si. materials were demonstrated as MCM-41, which has mild Brønsted acid
Though single alkali desilication is an effective and promising way to strength. The results of hydroisomerization of n-dodecane showed that
produce hierarchical pore structure both in the lab and in industry, it catalysts using ZSM-22 with MCM-41 mesoporous structure as acid
does have some disadvantages. For example, the limited mesopores support can evidently enhance the total isomers selectivity from 47.09%
introduction, ununiformed mesopores size, and destruction of micro­ to 91.70% at the conversion higher than 90%. This was attributed to the
pores [111]. In addition, a reduction of crystallinity was also found mesoporous MCM-41 on the lateral surface of ZSM-22 not only can in­
[112–115]. To overcome these negative effects caused by single alkali crease the pore mouth acid sites, which can lead to the increase of “key-
desilication in hydroisomerization of long-chain alkanes, modifications lock” catalysis, but also can adjust the overall acid strength. Similar
have been made based on protective strategy or recrystallization positive effect caused by recrystallization on enhancing textual prop­
strategy. erties during desilication was also reported on ZSM-5 [119].
Zhang et al. [116] reported the HDA-protected desilication of high From above summarized studies, it can be seen that the mesopores
Si/Al ratio ZSM-48 zeolites. In their experiments, the desilication of fabricated by “top down” method can evidently enhance the isomers
homemade ZSM-48 zeolites with a high Si/Al ratio equals to 200 was selectivity. It is usually attributed to the shortened diffusion path and
studied using NaOH solution or NaOH+ HDA solution. Though all the enhanced pore mouth exposure. However, the disadvantages caused by
samples showed desilication of ZSM-48 zeolites caused crystallinity single alkali desilication is very common. Therefore, to avoid the
reduction, the proper ratio of NaOH and HDA in the solution exhibited negative effects of single alkali desilication, strategies such as protective
the best preservation of crystallinity due to the protective effect of HDA. desilication and simultaneous desilication and recrystallization should
So, in the hydroisomerization of n-C16, as-synthesized hierarchical Ni/ be considered and utilized. And from the aspect of environmental pro­
ZSM-48 catalysts with well-preserved crystallinity showed the highest tection, the species of alkali used in desilication should be further
isomers selectivity. Wang et al. [117] reported the synthesis of hierar­ modified to avoid the pollution caused by broadly used NaOH.
chical ZSM-22 with a dual-protected alkali post-treatment strategy. To
preserve the crystallinity of ZSM-22 and maintain high desilication ef­ 4.2. Effect of crystal size
ficiency [111] at the same time, they combined partial detemplation and
CTAB protection in alkali treatment. This unique strategy resulted in It is known that the process of alkene intermediates isomerizes on the
mesopores formation centered at 3–4 nm. As shown in Fig. 6, in addition active protonic sites and the pathway length between two metal sites
to their uniformity, more pore mouth exposure caused by the protective play an important role in hydroisomerization of n-alkanes. Since
effect was confirmed. The protection effect of partial detemplation was bifunctional catalysts with smaller crystal size allows more accessibility
attributed to the template can prevent the contact between OH− 1 and Si- of active protonic sites and shortened pathway length of alkene in­
O-Si linkages inside the micropores, therefore, desilication can only termediates between two metal sites, downsizing the crystal is regarded
happen in the template free region. While CATB micelles can stabilize as a promising route to attain higher selectivity of isomers.

12
C. Wei et al. Fuel xxx (xxxx) xxx

Zhang et al. [120] studied the effect of the journey distance of alkene catalysts are the same in principle for the both parts. There are 3 stages
intermediates within the micropores, and demonstrated that zeolite of reaction scheme according to the nMe/nA ratio. For both light alkanes
particle size has a great influence on the final selectivity of isomers in and long-chain alkanes, the nMe/nA ratio should reach to the third stage
hydroisomerization. They synthesized micro/mesoporous Y/MCM-41 to guarantee the activity and stability of catalysts, as well as to lower the
composites, comprised of mesoporous MCM-41 material on the outer tendency of cracking, especially for long-chain alkanes. However, due to
layer and the inner alkali treated Y zeolites with different crystal size. the difference of reactivity between light alkanes and long-chain al­
The MCM-41 material on the outer layer on one hand is the main support kanes, the exact number of nMe/nA ratio are not the same. Besides the
of Pt metal particles, on the other hand, it can prevent the aggregation of nMe/nA ratio, the acid strength of their catalysts’ acid supports also has
Y zeolite. Therefore, on the Pt loaded catalysts, n-alkene intermediates significant difference. For light alkanes, people tend to use zeolites with
mainly formed on MCM-41, and the successive isomerization happened strong acid strength. While for long-chain alkanes, emphasis is always
inside inner Y zeolites. In the n-C12 hydroisomerization test of Pt loaded placed on supports with weaker acid strength. General explanation is the
catalysts with similar metal–acid balance but evidently different average high cracking tendency and weak diffusivity of the reaction in­
crystal size of Y zeolites, namely 653 nm (Pt/MY-0), 548 nm (Pt/MY-2), termediates of long-chain alkanes. The intermediates will undergo
421 nm (Pt/MY-5), respectively, the resulted selectivity of i-C12 excessive isomerization and successive cracking if the acid strength is
exhibited significant difference. It is found that the smaller crystal size of strong. However, there exists a dispute. Noh et al. argues the weak acid
Y zeolites in the catalysts were, the higher the selectivity was. And this strength will increase the intrinsic β-scission/isomerization rate ratios.
result is valid in each conversion level. This was attributed to the In addition, for long-chain alkanes, the so called “pore mouth” and “key
cracking reaction was suppressed by shortened diffusion pathlength and lock” mechanism plays an important role in their hydroisomerization.
reduced active acid sites the n-alkene intermediates will encounter be­ Therefore, efforts have been made to tune the acid sites location to attain
tween two metal sites. higher isomer selectivity. Considering the weak diffusivity of long-chain
Li et al. [121] reported the hydroisomerization of n-Decane using Pt/ alkanes, researches have been made on the modifications of the textural
Y catalysts with different crystal size. Alkali treatment was used to structure of zeolites and molecular sieves based catalysts. Emphasis are
downsize zeolite Y. Pt loaded catalysts based on alkali treated and mainly placed on two directions, i.e., fabrication of hierarchical pore
originate Y were designated as Pt/AY-akl and Pt/AY-ori, respectively. structure and downsizing of crystal size. In both ways, the enhancement
The average particle size of alkali treated Y zeolites is 501 nm, which is of isomer yields is remarkable, indicating more attention should be paid
much smaller than of the origin Y zeolites with 687 nm. The hydro­ on their textural properties.
isomerization of n-Decane on Pt/AY-akl and Pt/AY-ori showed that In the future, hydroisomerization will still play an important role in
catalyst with smaller crystal size exhibited higher isomer yield by about fuel and lubricants processing. However, whether for light alkanes or
10%. It was attributed to the higher resistance to multiple branching and long-chain alkanes, people should concentrate on the following aspects
skeletal cracking of Pt/Y-alk. Smaller crystal size of Y zeolites reduced to solve the emerging problems or optimize the overall process:
the probability of olefinic intermediates encounter more active protonic
acid sites inside the micropores and relieved the diffusion resistance. 1. The high cost caused by loading noble metals. Efforts should be paid
Therefore, the selectivity of isomers was enhanced by suppressing the on enhancing the hydrogenation efficiency of noble metals, and
undesired secondary cracking. finally the loading amount can be reduced. Another promising di­
Ge et al. [122] studied the effect of particle size in Pt/SAPO-11 rection is to partially or entirely replace it by applying a new form of
catalysts in hydroisomerization of n-dodecane. SAPO-11 molecular hydrogenation component.
sieves with different particle size were synthesized using modified dry- 2. The poor control of reaction degree. Since isomerization and
gel conversion. The particle size of as-synthesized SAPO-11 can be cracking of alkanes are consecutive reactions, the well control of
tuned from 65 nm to 450 nm by adjusting water content from 5% to reaction degree is a key to enhance isomers selectivity, especially for
50%. The analysis of Pt/SAPO-11 catalysts with different particle size long-chain alkanes. However, it’s still a great challenge to limit the
also showed the amount of protonic acid sites decreased with particle reaction degree to stop at isomerization. The proper coupling of
size. While the hydroisomerization of n-dodecane showed one can attain separation unit may help.
the highest isomer yields of ~75% on catalyst 1.4PtNS with smallest 3. Lack of control over the precise structure of the product. The struc­
crystal size. It was attributed to the reduced crystal size not only reduces tures of alkane isomers are closely related to their properties. For
the nH+ between two Pt sites, but also reduces the resistance time of gasoline distillates, highly branched isomers are favored for their
alkene intermediates within micropores. And the reduced nH+ between high octane number. However, for diesel distillates, terminal mono-
two Pt sites and resistance avoid cracking reaction. methyl-branched isomers are ideal due to their high cetane number
The studies introduced above suggest that the downsizing of acid and outstanding low temperature flow properties. For lubricants, the
support crystal size is an effective way to enhance the isomers selec­ pour point and viscosity index both decrease as the branching posi­
tivity. The enhanced selectivity is always attributed to the shortened tion is closer to the center of the straight chain. A balance between
pathway for reaction intermediates; thus, the intermediates may these two properties should be considered according to situation of
encounter fewer acid sites during the diffusion between two metallic practice application, with adjusting the isomerization degree and
sites. However, the approach to reduce size of acid supports should be position of side chains of the products. However, in practice, no
further developed, considering the economic and environmental matter which fraction is isomerized, the product structure varies in
aspects. branching degree and side chain position. Therefore, it’s of great
importance to optimize the control of isomers structure to enhance
5. Conclusion and outlook the quality of products.

In this review, we compared the hydroisomerization of linear alkanes Declaration of Competing Interest
of C5-C6 and long-chain ones. In the view of their mechanism, we can
find that the isomerization of alkene intermediates of long-chain alkanes The authors declare that they have no known competing financial
is more complicated. For light alkanes, undesired β-scission can only go interests or personal relationships that could have appeared to influence
through type C and type D mechanism, which are much slower than the the work reported in this paper.
type A and type B cracking of long-chain alkanes. This indicates long-
chain alkanes have stronger tendency of cracking, and it is consistent
with the lower isomer yields. The effects of metal acid balance of the

13
C. Wei et al. Fuel xxx (xxxx) xxx

Acknowledgment [24] Yang Ye, Xu Lu, Lyu Y, Liu X, Yan Z. Controllable synthesis of SAPO-11/5
intergrowth zeolite for hydroisomerization of n-hexane. Microporous Mesoporous
Mater 2021;313:110857. https://doi.org/10.1016/j.micromeso.2020.110857.
This work was supported by National Natural Science Foundation of [25] Tamizhdurai P, Ramesh A, Krishnan PS, Narayanan S, Shanthi K, Sivasanker S.
China (No. 22021004). Effect of acidity and porosity changes of dealuminated mordenite on n-pentane,
n-hexane and light naphtha isomerization. Microporous Mesoporous Mater 2019;
287:192–202. https://doi.org/10.1016/j.micromeso.2019.06.012.
References [26] Li W, Chi K, Liu H, Ma H, Qu W, Wang C, et al. Skeletal isomerization of n
-pentane: a comparative study on catalytic properties of Pt/WOx –ZrO2 and Pt/
[1] Liu S, Zhang L, Zhang L, Zhang H, Ren J. Function of well-established mesoporous ZSM-22. Appl Catal A 2017;537:59–65. https://doi.org/10.1016/j.
layers of recrystallized ZSM-22 zeolites in the catalytic performance of n-alkane apcata.2017.03.005.
isomerization. New J Chem 2020;44(12):4744–54. https://doi.org/10.1039/ [27] Yu G, Chen X, Xue W, Ge L, Wang T, Qiu M, et al. Melting-assisted solvent-free
C9NJ06273D. synthesis of SAPO-11 for improving the hydroisomerization performance of n-
[2] Shi Y, Zhang J, Xing E, Xie Y, Cao H. Selective production of jet-fuel-range dodecane. Chin J Catal 2020;41(4):622–30.
alkanes from palmitic acid over Ni/H-MCM-49 with two independent pore [28] Kang Y, Rao X, Yuan P, Wang C, Wang T, Yue Y. Al-functionalized mesoporous
systems. Ind Eng Chem Res 2019;58(47):21341–9. https://doi.org/10.1021/acs. SBA-15 with enhanced acidity for hydroisomerization of n-octane. Fuel Process
iecr.9b04130. Technol 2021;215:106765. https://doi.org/10.1016/j.fuproc.2021.106765.
[3] Rodionova LI, Knyazeva EE, Konnov SV, Ivanova II. Application of nanosized [29] Guo C, Wang W, Zhang Y, Lin H, Jia G, Li T, et al. Influences of the metal-acid
zeolites in petroleum chemistry: synthesis and catalytic properties (review). Pet proximity of Pd-SAPO-31 bifunctional catalysts for n-hexadecane
Chem 2019;59(4):455–70. https://doi.org/10.1134/S0965544119040133. hydroisomerization. Fuel Process Technol 2021;214:106717. https://doi.org/
[4] Song H, Zhao L, Wang N. Rare earth metals modified Ni–S2O82− /ZrO2–Al2O3 10.1016/j.fuproc.2020.106717.
catalysts for n-pentane isomerization. Chin J Chem Eng 2017;25(1):74–8. [30] Coonradt HL, Garwood WE. Mechanism of hydrocracking. Reactions of paraffins
https://doi.org/10.1016/j.cjche.2016.06.004. and olefins. Ind Eng Chem Process Des Dev 1964;3(1):38–45. https://doi.org/
[5] Liu P, Yao Y, Zhang X, Wang J. Rare Earth metals ion-exchanged β-zeolites as 10.1021/i260009a010.
supports of platinum catalysts for hydroisomerization of n-heptane. Chin J Chem [31] Claude MC, Martens JA. Monomethyl-Branching of long n-alkanes in the range
Eng 2011;19(2):278–84. https://doi.org/10.1016/s1004-9541(11)60166-3. from decane to tetracosane on Pt/H-ZSM-22 bifunctional catalyst. J Catal 2000;
[6] Qin B, Zhang X, Zhang Z, Ling F, Sun W. Synthesis, characterization and catalytic 190(1):39–48. https://doi.org/10.1006/jcat.1999.2714.
properties of Y-β zeolite composites. Pet Sci 2011;8(2):224–8. https://doi.org/ [32] Martens JA, Parton R, Uytterhoeven L, Jacobs PA, Froment GF. Selective
10.1007/s12182-011-0139-8. conversion of decane into branched isomers: a comparison of platinum/ZSM-22,
[7] Chi K, Zhao Z, Tian Z, Hu S, Yan L, Li T, et al. Hydroisomerization performance of platinum/ZSM-5 and platinum/USY zeolite catalyst. Appl Catal 1991;76(1):
platinum supported on ZSM-22/ZSM-23 intergrowth zeolite catalyst. Pet Sci 95–116. https://doi.org/10.1016/0166-9834(91)80007-J.
2013;10(2):242–50. https://doi.org/10.1007/s12182-013-0273-6. [33] Martens JA, Souverijns W, Verrelst W, Parton R, Froment GF, Jacobs PA. Selective
[8] Smeeth M, Spikes H, Gunsel S. Boundary film formation by viscosity index isomerization of hydrocarbon chains on external surfaces of zeolite crystals.
improvers. Tribol Trans 1996;39(3):726–34. https://doi.org/10.1080/ Angew Chem Int Ed Engl 1995;34(22):2528–30. https://doi.org/10.1002/
10402009608983590. anie.199525281.
[9] Denis J. The relationships between structure and rheological properties of [34] Forestière A, Olivier-Bourbigou H, Saussine L. Oligomerization of monoolefins by
hydrocarbons and oxygenated compounds used as base stocks. J Synth Lubr homogeneous catalysts. Oil Gas Sci Technol – Rev IFP 2009;64(6):649–67.
1984;1(3):201–38. https://doi.org/10.1002/jsl.3000010304. https://doi.org/10.2516/ogst/2009027.
[10] Jaroszewska K, Masalska A, Grzechowiak JR. Hydroisomerization of long-chain [35] Wang W, Liu C-J, Wu W. Bifunctional catalysts for the hydroisomerization of n-
bio-derived n-alkanes into monobranched high cetane isomers via a dual- alkanes: the effects of metal–acid balance and textural structure. Catal Sci
component catalyst bed. Fuel 2020;268:117239. https://doi.org/10.1016/j. Technol 2019;9(16):4162–87. https://doi.org/10.1039/c9cy00499h.
fuel.2020.117239. [36] Cheng K, Wal LI, Yoshida H, Oenema J, Harmel J, Zhang Z, et al. Impact of the
[11] Sazama P, Pastvova J, Kaucky D, Moravkova J, Rathousky J, Jakubec I, et al. Does spatial organization of bifunctional metal-zeolite catalysts on the
hierarchical structure affect the shape selectivity of zeolites? Example of hydroisomerization of light alkanes. Angew Chem Int Ed Engl 2020;59(9):
transformation of n-hexane in hydroisomerization. J Catal 2018;(364):262–70. 3592–600. https://doi.org/10.1002/anie.201915080.
https://doi.org/10.1016/j.jcat.2018.05.010. [37] Liu L, Zhang M, Wang Li, Zhang X, Li G. Construction of ordered mesopores
[12] Pastvova J, Pilar R, Moravkova J, Kaucky D, Rathousky J, Sklenak S, et al. outside MTT zeolite for efficient hydroisomerization. Appl Catal A 2020;602:
Tailoring the structure and acid site accessibility of mordenite zeolite for 117664. https://doi.org/10.1016/j.apcata.2020.117664.
hydroisomerisation of n-hexane. Appl Catal A 2018;562:159–72. https://doi.org/ [38] Zhao X, Liu W, Wang J, Yang W, Zhu X, Zhu K. Interface mediated crystallization
10.1016/j.apcata.2018.05.035. of plate-like SAPO-41 crystals to promote catalytic hydroisomerization. Appl
[13] Triwahyono S, Jalil AA, Izan SM, Jamari NS, Fatah NAA. Isomerization of linear Catal A 2020;602:117738. https://doi.org/10.1016/j.apcata.2020.117738.
C5–C7 over Pt loaded on protonated fibrous silica@Y zeolite (Pt/HSi@Y). [39] Smit B, Maesen TLM. Molecular simulations of zeolites: adsorption, diffusion, and
J Energy Chem 2019;37:163–71. https://doi.org/10.1016/j.jechem.2019.02.016. shape selectivity. Chem Rev 2008;108(10):4125–84. https://doi.org/10.1021/
[14] Lee W, Yang CC, Cheng S. Isomerization of n-pentane over platinum promoted cr8002642.
tungstated zirconia supported on mesoporous SBA-15 prepared by supercritical [40] Alvarez F, Ribeiro FR, Perot G, Thomazeau C, Guisnet M. Hydroisomerization and
impregnation. J Chin Chem Soc 2020. https://doi.org/10.1002/jccs.202000473. hydrocracking of alkanes: 7. Influence of the balance between acid and
[15] Vera-Iturriaga J, Hernández-Pichardo ML, Montoya de la Fuente JA, Del Angel P, hydrogenating functions on the transformation ofn-decane on PtHY catalysts.
Gomora-Herrera D, Palacios-González E, et al. Hydroisomerization of n-hexane J Catal 1996;162(2):179–89. https://doi.org/10.1006/jcat.1996.0275.
over Pt/WOx-ZrO2-TiO2 catalysts. Catal Today 2021;360:12–9. https://doi.org/ [41] Rey J, Raybaud P, Chizallet C, Bučko T. Competition of secondary versus tertiary
10.1016/j.cattod.2019.08.043. carbenium routes for the Type B isomerization of alkenes over acid zeolites
[16] Wang P, Yue Y, Wang T, Bao X. Alkane isomerization over sulfated zirconia solid quantified by Ab initio molecular dynamics simulations. ACS Catal 2019;9(11):
acid system. Int J Energy Res 2020;44(5):3270–94. https://doi.org/10.1002/ 9813–28. https://doi.org/10.1021/acscatal.9b02856.
er.4995. [42] Chang CD, Prins R, Sinfelt JH, von Ballmoos R, Harris DH, Magee JS, et al.
[17] Wen C, Xu J, Wang X, Fan Yu. n-Heptane Hydroisomerization over a SO42–/ Energy-related catalysis: sections 3.7.6 – 3.12.4.2. In: Handbook of
ZrO2@SAPO-11 Composite-Based Catalyst Derived from the Growth of UiO-66 on Heterogeneous Catalysis; 1997. p. 1900–2006. https://doi.org/10.1002/
SAPO-11. Energy Fuels 2020;34(8):9498–508. https://doi.org/10.1021/acs. 9783527619474.ch14b.
energyfuels.0c01634. [43] Kim J, Han SW, Kim J-C, Ryoo R. Supporting nickel to replace platinum on zeolite
[18] Marcilly C. Present status and future trends in catalysis for refining and nanosponges for catalytic hydroisomerization of n-dodecane. ACS Catal 2018;8
petrochemicals. J Catal 2003;216(1–2):47–62. https://doi.org/10.1016/s0021- (11):10545–54. https://doi.org/10.1021/acscatal.8b03301.
9517(02)00129-x. [44] Lyu Y, Yu Z, Yang Ye, Liu Y, Zhao X, Liu X, et al. Metal and acid sites
[19] Primo A, Garcia H. Zeolites as catalysts in oil refining. Chem Soc Rev 2014;43 instantaneously prepared over Ni/SAPO-11 bifunctional catalyst. J Catal 2019;
(22):7548–61. https://doi.org/10.1039/C3CS60394F. 374:208–16. https://doi.org/10.1016/j.jcat.2019.04.031.
[20] Zschiesche C, Himsl D, Rakoczy R, Reitzmann A, Freiding J, Wilde N, et al. [45] Moreau C, Geneste P. In: Factors Affecting the Reactivity of Organic Model
Hydroisomerization of long-chain n-alkanes over bifunctional zeolites with 10- Compounds in Hydrotreating Reactions. Theoretical Aspects of Heterogeneous
membered- and 12-membered-ring pores. Chem Eng Technol 2018;41(1): Catalysis. Dordrecht: Springer Netherlands; 1990. p. 256–310. https://doi.org/
199–204. https://doi.org/10.1002/ceat.201700236. 10.1007/978-94-010-9882-3_7.
[21] Smit B, Maesen TL. Towards a molecular understanding of shape selectivity. [46] Weitkamp J, Jacobs PA, Martens JA. Isomerization and hydrocracking of C9
Nature 2008;451(7179):671–8. https://doi.org/10.1038/nature06552. through C16 n-alkanes on Pt/HZSM-5 zeolite. Appl Catal 1983;8(1):123–41.
[22] Busca G. Acid catalysts in industrial hydrocarbon chemistry. Chem Rev 2007;107 [47] Martens JA, Jacobs PA, Weitkamp J. Attempts to rationalize the distribution of
(11):5366–410. https://doi.org/10.1021/cr068042e. hydrocracked products. I qualitative description of the primary hydrocracking
[23] Aitani A, Akhtar MN, Al-Khattaf S, Jin Y, Koseoglo O, Klein MT. Catalytic modes of long chain paraffins in open zeolites. Appl Catal 1986;20(1):239–81.
upgrading of light naphtha to gasoline blending components: a mini review. https://doi.org/10.1016/0166-9834(86)80020-3.
Energy Fuels 2019;33(5):3828–43. https://doi.org/10.1021/acs. [48] Alvarez F, Ribeiro FR, Giannetto G, Chevalier F, Perot G, Guisnet M.
energyfuels.9b00704. Hydroisomerization and Hydrocracking of Alkanes. 5. hydroisomerization and
hydrocracking of N-hexane and N-heptane on PtHY catalysts. Effect of the
distribution of metallic and ACID sites. In: Jacobs PA, van Santen RA, editors.

14
C. Wei et al. Fuel xxx (xxxx) xxx

Studies in Surface Science and Catalysis. Elsevier; 1989. p. 1339–48. https://doi. [73] Lv G, Wang C, Chi K, Liu H, Wang P, Ma H, et al. Effects of Pt site distributions on
org/10.1016/S0167-2991(08)62018-2. the catalytic performance of Pt/SAPO-11 for n-dodecane hydroisomerization.
[49] Lyu Y, Yu Z, Yang Ye, Wang X, Zhao X, Liu X, et al. Metal-acid balance in the in- Catal Today 2018;316:43–50. https://doi.org/10.1016/j.cattod.2018.04.072.
situ solid synthesized Ni/SAPO-11 catalyst for n-hexane hydroisomerization. Fuel [74] Smirnova MY, Kikhtyanin OV, Smirnov MY, Kalinkin AV, Titkov AI, Ayupov AB,
2019;243:398–405. https://doi.org/10.1016/j.fuel.2019.01.013. et al. Effect of calcination temperature on the properties of Pt/SAPO-31 catalyst
[50] Giannetto GE, Perot GR, Guisnet MR. Hydroisomerization and hydrocracking of in one-stage transformation of sunflower oil to green diesel. Appl Catal A 2015;
n-alkanes. 1. Ideal hydroisomerization PtHY catalysts. Ind Eng Chem Prod Res 505:524–31. https://doi.org/10.1016/j.apcata.2015.06.019.
Dev 1986;25(3):481–90. https://doi.org/10.1021/i300023a021. [75] Shakun AN, Fedorova ML. Isomerization of light gasoline fractions: the efficiency
[51] Batalha N, Pinard L, Bouchy C, Guillon E, Guisnet M. n-Hexadecane of different catalysts and technologies. Catal Ind 2014;6(4):298–306. https://doi.
hydroisomerization over Pt-HBEA catalysts. Quantification and effect of the org/10.1134/S2070050414040163.
intimacy between metal and protonic sites. J Catal 2013;307:122–31. https://doi. [76] Sie ST, Maxwell JE, Minderhoud JK, Stork WJH, van Veen JAR, Weitkamp J, et al.
org/10.1016/j.jcat.2013.07.014. Energy-related catalysis: sections 3.12.5 – 3.19.8. Handbook of heterogeneous.
[52] Lucas Ad, Ramos MJ, Dorado F, Sánchez P, Valverde JL. Influence of the Si/Al Catalysis 1997:2006–122. https://doi.org/10.1002/9783527619474.ch14c.
ratio in the hydroisomerization of n-octane over platinum and palladium beta [77] Choudhury IR, Hayasaka K, Thybaut JW, Laxmi Narasimhan CS, Denayer JF,
zeolite-based catalysts with or without binder. Appl Catal A 2005;289(2):205–13. Martens JA, et al. Pt/H-ZSM-22 hydroisomerization catalysts optimization guided
https://doi.org/10.1016/j.apcata.2005.05.001. by Single-Event MicroKinetic modeling. J Catal 2012;290:165–76. https://doi.
[53] Ali A-G-A, Ali LI, Aboul-Fotouh SM, Hydroisomerization A-G. hydrocracking and org/10.1016/j.jcat.2012.03.015.
dehydrocyclization of n-pentane and n-hexane using mono- and bimetallic [78] Laxmi Narasimhan CS, Thybaut JW, Marin GB, Jacobs PA, Martens JA,
catalysts promoted with fluorine. Appl Catal A 2001;215(1):161–73. https://doi. Denayer JF, et al. Kinetic modeling of pore mouth catalysis in the
org/10.1016/S0926-860X(01)00515-4. hydroconversion of n-octane on Pt-H-ZSM-22. J Catal 2003;220(2):399–413.
[54] Tian S, Chen J. Hydroisomerization of n-dodecane on a new kind of bifunctional [79] Wang X, Zhang X, Wang Q. N-dodecane hydroisomerization over Pt/ZSM-22:
catalyst: Nickel phosphide supported on SAPO-11 molecular sieve. Fuel Process controllable microporous Brönsted acidity distribution and shape-selectivity.
Technol 2014;122:120–8. https://doi.org/10.1016/j.fuproc.2014.01.031. Appl Catal A 2020;590:117335. https://doi.org/10.1016/j.apcata.2019.117335.
[55] Kinger G, Vinek H. n-Nonane hydroconversion on Ni and Pt containing HMFI, [80] Kim MY, Lee K, Choi M. Cooperative effects of secondary mesoporosity and acid
HMOR and HBEA. Appl Catal A 2001;218(1):139–49. https://doi.org/10.1016/ site location in Pt/SAPO-11 on n -dodecane hydroisomerization selectivity.
S0926-860X(01)00629-9. J Catal 2014;319:232–8. https://doi.org/10.1016/j.jcat.2014.09.001.
[56] Weitkamp J. Isomerization of long-chain n-alkanes on a Pt/CaY zeolite catalyst. [81] Niu P, Xi H, Ren J, Lin M, Wang Q, Chen X, et al. Micropore blocked core–shell
Ind Eng Chem Prod Res Dev 1982;21(4):550–8. https://doi.org/10.1021/ ZSM-22 designed via epitaxial growth with enhanced shape selectivity and high
i300008a008. n-dodecane hydroisomerization performance. Catal Sci Technol 2018;8(24):
[57] Weitkamp J. In: The influence of chain length in hydrocracking and 6407–19.
hydroisomerization of n-alkanes. hydrocracking and hydrotreating. American [82] Lv G, Wang C, Wang P, Sun L, Liu H, Qu W, et al. Pt/ZSM-22 with partially filled
Chemical Society; 1975. p. 1–27. micropore channels as excellent shape-selective hydroisomerization catalyst.
[58] Viswanadham N, Dixit L, Gupta JK, Garg MO. Effect of acidity and porosity ChemCatChem 2019;11(5):1431–6. https://doi.org/10.1002/cctc.201801695.
changes of dealuminated mordenites on n-hexane isomerization. J Mol Catal A: [83] Zhang F, Liu Y, Sun Qi, Dai Z, Gies H, Wu Q, et al. Design and preparation of
Chem 2006;258(1–2):15–21. https://doi.org/10.1016/j.molcata.2006.04.067. efficient hydroisomerization catalysts by the formation of stable SAPO-11
[59] Saxena SK, Kamble R, Singh M, Garg MO, Viswanadham N. Effect of acid molecular sieve nanosheets with 10–20 nm thickness and partially blocked acidic
treatments on physico-chemical properties and isomerization activity of sites. Chem Commun (Camb) 2017;53(36):4942–5.
mordenite. Catal Today 2009;141(1–2):215–9. https://doi.org/10.1016/j. [84] Weisz PB. Polyfunctional Heterogeneous Catalysis. In: Eley DD, Selwood PW,
cattod.2008.03.013. Weisz PB, Balandin AA, De Boer JH, Debye PJ, editors. Advances in Catalysis.
[60] Chen Z, Liu S, Wang H, Ning Q, Zhang H, Yun Y, et al. Synthesis and Academic Press; 1962. p. 137–90.
characterization of bundle-shaped ZSM-22 zeolite via the oriented fusion of [85] Zecevic J, Vanbutsele G, de Jong KP, Martens JA. Nanoscale intimacy in
nanorods and its enhanced isomerization performance. J Catal 2018;361:177–85. bifunctional catalysts for selective conversion of hydrocarbons. Nature 2015;528
https://doi.org/10.1016/j.jcat.2018.02.019. (7581):245–8. https://doi.org/10.1038/nature16173.
[61] Parmar S, Pant KK, John M, Kumar K, Pai SM, Newalkar BL. Hydroisomerization [86] Ben Moussa O, Tinat L, Jin X, Baaziz W, Durupthy O, Sayag C, et al.
of n-hexadecane over Pt/ZSM-22 framework: effect of divalent cation exchange. Heteroaggregation and selective deposition for the fine design of
J Mol Catal A: Chem 2015;404–405:47–56. https://doi.org/10.1016/j. nanoarchitectured bifunctional catalysts: application to hydroisomerization. ACS
molcata.2015.04.012. Catal 2018;8(7):6071–8. https://doi.org/10.1021/acscatal.8b01461.
[62] Girgis MJ, Tsao YP. Impact of catalyst metal− acid balance in n-hexadecane [87] Oenema J, Harmel J, Vélez RP, Meijerink MJ, Eijsvogel W, Poursaeidesfahani A,
hydroisomerization and hydrocracking. Ind Eng Chem Res 1996;35(2):386–96. et al. Influence of nanoscale intimacy and zeolite micropore size on the
https://doi.org/10.1021/ie9501586. performance of bifunctional catalysts for n-heptane hydroisomerization. ACS
[63] Feng Z, Wang W, Wang Yu, Bai X, Su X, Yang L, et al. Hydroisomerization of n- Catal 2020;10(23):14245–57. https://doi.org/10.1021/acscatal.0c03138.
decane over the Pd/ZSM-22 bifunctional catalysts: the effects of dynamic and [88] Samad JE, Blanchard J, Sayag C, Louis C, Regalbuto JR. The controlled synthesis
static crystallization to the zeolite. Microporous Mesoporous Mater 2019;274: of metal-acid bifunctional catalysts: the effect of metal:acid ratio and metal-acid
1–8. https://doi.org/10.1016/j.micromeso.2018.07.012. proximity in Pt silica-alumina catalysts for n-heptane isomerization. J Catal 2016;
[64] Niu P, Xi H, Ren J, Lin M, Wang Q, Jia L, et al. High selectivity for n-dodecane 342:203–12. https://doi.org/10.1016/j.jcat.2016.08.004.
hydroisomerization over highly siliceous ZSM-22 with low Pt loading. Catal Sci [89] Mendes PSF, Silva JM, Ribeiro MF, Daudin A, Bouchy C. Bifunctional intimacy
Technol 2017;7(21):5055–68. and its interplay with metal-acid balance in shaped hydroisomerization catalysts.
[65] Liu S, Ren J, Zhu S, Zhang H, Lv E, Xu J, et al. Synthesis and characterization of ChemCatChem 2020;12(18):4582–92. https://doi.org/10.1002/cctc.202000624.
the Fe-substituted ZSM-22 zeolite catalyst with high n-dodecane isomerization [90] Hu C, Sun J, Kang D, Zhu Q, Yang Y. Mechanistic insights into complete
performance. J Catal 2015;330:485–96. https://doi.org/10.1016/j. hydrogenation of 1,3-butadiene over Pt/SiO2: effect of Pt dispersion and kinetic
jcat.2015.07.027. analysis. Catal Sci Technol 2017;7(13):2717–28. https://doi.org/10.1039/
[66] Noh G, Zones SI, Iglesia E. Consequences of acid strength and diffusional C7CY00534B.
constraints for alkane isomerization and β-scission turnover rates and selectivities [91] Liu L, Corma A. Metal catalysts for heterogeneous catalysis: from single atoms to
on bifunctional metal-acid catalysts. J Phys Chem C 2018;122(44):25475–97. nanoclusters and nanoparticles. Chem Rev 2018;118(10):4981–5079. https://doi.
https://doi.org/10.1021/acs.jpcc.8b08460. org/10.1021/acs.chemrev.7b00776.
[67] Sánchez P, Dorado F, Ramos MJ, Romero R, Jiménez V, Valverde JL. [92] Wang Y, Tao Z, Wu B, Chen H, Xu J, Yang Y, et al. Shape-controlled synthesis of
Hydroisomerization of C6–C8 n-alkanes, cyclohexane and benzene over Pt particles and their catalytic performances in the n -hexadecane
palladium and platinum beta catalysts agglomerated with bentonite. Appl Catal A hydroconversion. Catal Today 2016;259:331–9. https://doi.org/10.1016/j.
2006;314(2):248–55. https://doi.org/10.1016/j.apcata.2006.08.025. cattod.2015.06.017.
[68] Chica A, Corma A. Hydroisomerization of pentane, hexane, and heptane for [93] Wang D, Liu J, Cheng X, Kang X, Wu A, Tian C, et al. Trace Pt clusters dispersed
improving the octane number of gasoline. J Catal 1999;187(1):167–76. on SAPO-11 promoting the synergy of metal sites with acid sites for high-effective
[69] Denayer JF, Baron GV, Jacobs PA, Martens JA. Competitive physisorption effects hydroisomerization of n-alkanes. small. Methods 2019;3(5):1800510. https://doi.
in hydroisomerisation of n-alkane mixtures on Pt/Y and Pt/USY zeolite catalysts. org/10.1002/smtd.201800510.
Phys Chem Chem Phys 2000;2(5):1007–14. https://doi.org/10.1039/A907619K. [94] Geerts L, Ramachandran RK, Dendooven J, Radhakrishnan S, Seo JW,
[70] Roldán R, Romero FJ, Jiménez-Sanchidrián C, Marinas JM, Gómez JP. Influence Detavernier C, et al. Creation of gallium acid and platinum metal sites in
of acidity and pore geometry on the product distribution in the bifunctional zeolite hydroisomerization and hydrocracking catalysts by atomic
hydroisomerization of light paraffins on zeolites. Appl Catal A 2005;288(1): layer deposition. Catal Sci Technol 2020;10(6):1778–88.
104–15. https://doi.org/10.1016/j.apcata.2005.04.029. [95] Al-Rawi UA, Sher F, Hazafa A, Rasheed T, Al-Shara NK, Lima EC, et al. Catalytic
[71] Jobic H, Theodorou DN. Diffusion of long n-alkanes in silicalite. A comparison activity of pt loaded zeolites for hydroisomerization of n-hexane using
between neutron scattering experiments and hierarchical simulation results. supercritical CO2. Ind Eng Chem Res 2020;59(51):22092–106. https://doi.org/
J Phys Chem B 2006;110(5):1964–7. https://doi.org/10.1021/jp056924w. 10.1021/acs.iecr.0c05184.
[72] Jobic H, Theodorou DN. Quasi-elastic neutron scattering and molecular dynamics [96] Wang D, Kang X, Gu Y, Zhang H, Liu J, Wu A, et al. Electronic tuning of Ni by Mo
simulation as complementary techniques for studying diffusion in zeolites. species for highly efficient hydroisomerization of n-alkanes comparable to Pt-
Microporous Mesoporous Mater 2007;102(1):21–50. https://doi.org/10.1016/j. based catalysts. ACS Catal 2020;10(18):10449–58. https://doi.org/10.1021/
micromeso.2006.12.034. acscatal.0c01159.

15
C. Wei et al. Fuel xxx (xxxx) xxx

[97] Chen H, Yi F, Ma C, Gao X, Liu S, Tao Z, et al. Hydroisomerization of n-heptane on [110] Wei X, Smirniotis PG. Development and characterization of mesoporosity in ZSM-
a new kind of bifunctional catalysts with palladium nanoparticles encapsulating 12 by desilication. Microporous Mesoporous Mater 2006;97(1–3):97–106.
inside zeolites. Fuel 2020;268:117241. https://doi.org/10.1016/j. https://doi.org/10.1016/j.micromeso.2006.01.024.
fuel.2020.117241. [111] Verboekend D, Chabaneix AM, Thomas K, Gilson J-P, Pérez-Ramírez J.
[98] Harmel J, Roberts T, Zhang Z, Sunley G, de Jongh P, de Jong KP. Bifunctional Mesoporous ZSM-22 zeolite obtained by desilication: peculiarities associated with
molybdenum oxide/acid catalysts for hydroisomerization of n-heptane. J Catal crystal morphology and aluminium distribution. CrystEngComm 2011;13(10):
2020;390:161–9. https://doi.org/10.1016/j.jcat.2020.08.004. 3408–16. https://doi.org/10.1039/C0CE00966K.
[99] Fan Y, Xiao H, Shi G, Liu H, Bao X. Alkylphosphonic acid- and small amine- [112] Groen JC, Bach T, Ziese U, Paulaime-van Donk AM, de Jong KP, Moulijn JA, et al.
templated synthesis of hierarchical silicoaluminophosphate molecular sieves with Creation of hollow zeolite architectures by controlled desilication of Al-Zoned
high isomerization selectivity to di-branched paraffins. J Catal 2012;285(1): ZSM-5 crystals. J Am Chem Soc 2005;127(31):10792–3. https://doi.org/
251–9. https://doi.org/10.1016/j.jcat.2011.09.037. 10.1021/ja052592x.
[100] Tao S, Li X, Lv G, Wang C, Xu R, Ma H, et al. Highly mesoporous SAPO-11 [113] Groen JC, Moulijn JA, Pérez-Ramírez J. Desilication: on the controlled generation
molecular sieves with tunable acidity: facile synthesis, formation mechanism and of mesoporosity in MFI zeolites. J Mater Chem 2006;16(22):2121–31. https://doi.
catalytic performance in hydroisomerization of n-dodecane. Catal Sci Technol org/10.1039/B517510K.
2017;7(23):5775–84. https://doi.org/10.1039/c7cy01819c. [114] Groen JC, Peffer LAA, Moulijn JA, Pérez-Ramírez J. On the introduction of
[101] Tao S, Li X, Wang X, Wei Y, Jia Y, Ju J, et al. Facile synthesis of hierarchical intracrystalline mesoporosity in zeolites upon desilication in alkaline medium.
nanosized single-crystal aluminophosphate molecular sieves from highly Microporous Mesoporous Mater 2004;69(1–2):29–34. https://doi.org/10.1016/j.
homogeneous and concentrated precursors. Angew Chem Int Ed 2020;59(9): micromeso.2004.01.002.
3455–9. https://doi.org/10.1002/anie.201915144. [115] Pérez-Ramírez J, Christensen CH, Egeblad K, Christensen CH, Groen JC.
[102] Wen C, Wang X, Xu J, Fan Yu. Hierarchical SAPO-11 molecular sieve-based Hierarchical zeolites: enhanced utilisation of microporous crystals in catalysis by
catalysts for enhancing the double-branched hydroisomerization of alkanes. Fuel advances in materials design. Chem Soc Rev 2008;37(11):2530. https://doi.org/
2019;255:115821. https://doi.org/10.1016/j.fuel.2019.115821. 10.1002/chin.200905229.
[103] Zhang P, Liu H, Yue Y, Zhu H, Bao X. Direct synthesis of hierarchical SAPO-11 [116] Zhang M, Li C, Chen X, Chen Y, Liang C. Hierarchical ZSM-48-supported nickel
molecular sieve with enhanced hydroisomerization performance. Fuel Process catalysts with enhanced hydroisomerization performance of hexadecane. Ind Eng
Technol 2018;179:72–85. https://doi.org/10.1016/j.fuproc.2018.06.012. Chem Res 2019;58(43):19855–61. https://doi.org/10.1021/acs.iecr.9b04415.
[104] Tao S, Li X, Gong H, Jiang Q, Yu W, Ma H, et al. Confined-space synthesis of [117] Wang X, Zhang X, Wang Q. n-Dodecane hydroisomerization over hierarchical
hierarchical MgAPO-11 molecular sieves with good hydroisomerization ZSM-22 prepared by a dual-protected alkali treatment. Ind Eng Chem Res 2019;58
performance. Microporous Mesoporous Mater 2018;262:182–90. https://doi.org/ (19):8495–505. https://doi.org/10.1021/acs.iecr.8b06450.
10.1016/j.micromeso.2017.11.041. [118] Liu S, Ren J, Zhang H, Lv E, Yang Y, Li Y-W. Synthesis, characterization and
[105] Jin D, Ye G, Zheng J, Yang W, Zhu K, Coppens M-O, et al. Hierarchical isomerization performance of micro/mesoporous materials based on H-ZSM-22
silicoaluminophosphate catalysts with enhanced hydroisomerization selectivity zeolite. J Catal 2016;335:11–23. https://doi.org/10.1016/j.jcat.2015.12.009.
by directing the orientated assembly of premanufactured building blocks. ACS [119] Yoo WC, Zhang X, Tsapatsis M, Stein A. Synthesis of mesoporous ZSM-5 zeolites
Catal 2017;7(9):5887–902. https://doi.org/10.1021/acscatal.7b01646. through desilication and re-assembly processes. Microporous Mesoporous Mater
[106] Jin D, Li L, Ye G, Ding H, Zhao X, Zhu K, et al. Manipulating the mesostructure of 2012;149(1):147–57. https://doi.org/10.1016/j.micromeso.2011.08.014.
silicoaluminophosphate SAPO-11 via tumbling-assisted, oriented assembly [120] Zhang Y, Liu D, Men Z, Huang Ke, Lv Y, Li M, et al. Hydroisomerization of n-
crystallization: a pathway to enhance selectivity in hydroisomerization. Catal Sci dodecane over bi-porous Pt-containing bifunctional catalysts: effects of alkene
Technol 2018;8(19):5044–61. https://doi.org/10.1039/c8cy01483c. intermediates’ journey distances within the zeolite micropores. Fuel 2019;236:
[107] Dudovich N, Oron D, Silberberg Y. Single-pulse coherently controlled nonlinear 428–36. https://doi.org/10.1016/j.fuel.2018.09.017.
Raman spectroscopy and microscopy. Nature 2002;418(6897):512–4. https://doi. [121] Li M, Zhang Y, Wang H, Yu S, Liu D, Wang Y. Influence of zeolite crystal size on
org/10.1038/nature00933. selective conversion of n-alkane: Controlling intermediates’ diffusion distances
[108] Triantafillidis CS, Vlessidis AG, Evmiridis NP. Dealuminated H− Y zeolites: inside the micropores. Fuel 2019;254:115709. https://doi.org/10.1016/j.
influence of the degree and the type of dealumination method on the structural fuel.2019.115709.
and acidic characteristics of H− Y zeolites. Ind Eng Chem Res 2000;39(2):307–19. [122] Ge L, Yu G, Chen X, Li W, Xue W, Qiu M, et al. Effects of particle size on
https://doi.org/10.1021/ie990568k. bifunctional Pt/SAPO-11 catalysts in the hydroisomerization of n-dodecane. New
[109] Hartmann M. Hierarchical zeolites: a proven strategy to combine shape selectivity J Chem 2020;44(7):2996–3003. https://doi.org/10.1039/c9nj06215g.
with efficient mass transport. Angew Chem Int Ed 2004;43(44):5880–2. https://
doi.org/10.1002/anie.200460644.

16

You might also like