Yaru Lina 2018
Yaru Lina 2018
Yaru Lina 2018
https://doi.org/10.1038/s41929-018-0078-5
The production of high-demand chemical commodities such as ethylene and propylene (methanol-to-olefins), hydrocarbons
(methanol-to-hydrocarbons), gasoline (methanol-to-gasoline) and aromatics (methanol-to-aromatics) from methanol—
obtainable from alternative feedstocks, such as carbon dioxide, biomass, waste or natural gas through the intermediate forma-
tion of synthesis gas—has been central to research in both academia and industry. Although discovered in the late 1970s, this
catalytic technology has only been industrially implemented over the past decade, with a number of large commercial plants
already operating in Asia. However, as is the case for other technologies, industrial maturity is not synonymous with full under-
standing. For this reason, research is still intense and a number of important discoveries have been reported over the last few
years. In this review, we summarize the most recent advances in mechanistic understanding—including direct C–C bond forma-
tion during the induction period and the promotional effect of zeolite topology and acidity on the alkene cycle—and correlate
these insights to practical aspects in terms of catalyst design and engineering.
I
n 1977, scientists at Mobile published the first patent on the meth- picture on the mechanism of reaction and to provide guidelines
anol-to-hydrocarbons (MTH) process1, claiming that “A lower for catalyst design. We believe that, despite a number of excellent
alcohol and/or ether feed is selectively converted to a mixture of reviews in this field, knowledge on mechanistic aspects and struc-
light olefins, including ethylene and propylene, by catalytic contact ture–property relationships in MTH has advanced tremendously
of the feed, for example methanol or dimethyl ether, at sub-atmo- over the last few years and we expect this Review to bridge the
spheric partial pressure, with certain crystalline aluminosilicate gap between those earlier reviews dedicated to mechanism ratio-
zeolite catalysts exemplified by ZSM-5” (Fig. 1a). The possibility nalization and catalyst engineering. To do so, we first summarize
to obtain olefins (methanol-to-olefins, MTO or DMTO if the feed- the recent efforts to unravel the mechanism that leads to the for-
stock is dimethylether), aromatics (methanol-to-aromatics, MTA) mation of the first C–C bond (that is, the coupling of two metha-
and/or gasoline (methanol-to-gasoline, MTG) without relying nol or dimethyl ether molecules), a topic of intense debate over the
on oil (methanol can be produced from a wide range of carbon- last few decades. Formation of the direct C–C bond continues with
containing sources, such as biomass, waste, coal, natural gas or even the formation of longer hydrocarbons inside the zeolite pores, this
carbon dioxide) makes this family of processes a convenient alter- mechanism along with the most recent spectroscopic evidences of
native to the classical production routes of high-demand chemicals its different steps are discussed in the second section. Finally, with
and intermediates. this mechanistic information in hand, we define a number of cata-
As it is the case for other established methods of the petrochemi- lyst design rules that should help the reader understand the effect
cal industry, such as fluid catalytic cracking (FCC)2, the MTH pro- of catalyst topology, acidity and reaction conditions on product dis-
cess — despite being already commercialized — still attracts a great tribution and catalyst deactivation. The Review is completed with
deal of attention from both industry and academia. Apart from the our personal view on future challenges both from the process and
obvious economic driving force in improving the process, from a fundamental point of view.
scientific point of view, understanding (and eventually controlling)
the complex reaction mechanism behind such a complex overall The direct C–C bond
stoichiometry remains a challenge. As we will elaborate below, the The MTH process involves a very complex reaction mechanism5–8,10.
process is dictated by a large set of elementary reactions and the full As of now, more than 20 different proposals have been postulated
picture of the MTH reaction mechanism is still elusive (Fig. 1b). On in the literature4. Surprisingly, the exact route for the formation of
one hand, the chosen reaction conditions are important, as not all the first carbon–carbon (C–C) bond during MTH was only very
reactions have the same kinetic order and energies of activation; on recently unveiled11–18. For a long time, the scientific community
the other hand, the zeolite microenvironment plays an even bigger assumed that the presence of traces of impurities (for example, in
role in defining product distribution and catalyst lifetime3. MTH the methanol, catalyst and/or carrier gas) was responsible for the
therefore serves as an outstanding example of complexity in zeolite formation of the direct C–C bond over any direct mechanism (that
chemistry and heterogeneous catalysis, as pinpointed in a number is, the coupling of two methanol molecules). This assumption was
of classical4 and more recent reviews5–9. primarily attributed to the lack of concrete experimental evidence
In this Review, we highlight current challenges of MTH chemis- in support of the direct mechanism19,20. Moreover, the feasibility
try by critically analysing the existing literature to give an updated of such direct coupling was anticipated to be low by theoreticians
King Abdullah University of Science and Technology, KAUST Catalysis Center, Advanced Catalytic Materials, Thuwal, Saudi Arabia. 2Inorganic Chemistry
1
and Catalysis, Debye Institute for Nanomaterials Science, Utrecht University, Utrecht, The Netherlands. 3These authors contributed equally: Irina Yarulina,
Abhishek Dutta Chowdhury. *e-mail: [email protected]; [email protected]
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
a Dessau ‘cracking’
mechanism
1983 1993 ‘Impurities as the 2006 2010 Direct C–C bond
MTH First MTH plant first C–C bond source’ Semi-commercial spectroscopic
discovery in New Zealand by Haw MTO plant in Belguim evidence
O Alkylation/
ring-cyclisation Arene Oligomerization
Coke
Zeolite CH3OR cycle
OR
CH3OH (ZeOH) (R=H, CH3)
CH3OCH3 –H2O
ZeOCH3 OR C2=-C5=
Initial C–C RO
Dehydration bond formation/ OCH3 Hydrogen transfer
Arene Alkanes
Methylation cycle
Directly formed
C–C bonds in bold Alkylation
Cracking
‘Direct’ mechanism Dual-cycle/HCP mechanism Deactivation
Fig. 1 | Milestones and mechanism development of the MTH process. a, MTH chronology shows that 40 years of combined efforts produced several
mechanistic concepts for both the initial induction and steady state periods. The current simplified mechanistic picture is shown in b. The simplified
scheme of the MTH mechanism reveals that although the aromatic cycle leads to the formation of light olefins, it also acts as a scaffold for coke formation;
thus, olefins formation is preferred over the olefins cycle (MTH, methanol-to-hydrocarbons; MTO, methanol-to-olefin; DMTO, dimethylether-to-olefin;
ZeOH, zeolite acid sites).
because of the high activation energies and unstable reaction inter- originated from the coupling between the carbene-like SMS and
mediates21,22. In 2006, Hunger and co-workers23 systematically methanol/dimethyl ether (DME) (Fig. 2b). Another convincing evi-
demonstrated that traces of organic impurities neither have any sig- dence in support of the existence of a carbene-type mechanism was
nificant influence on product distribution, nor do they control the provided recently by Weckhuysen et al.13, by employing 2D magic-
formation of hydrocarbon pool (HCP, vide infra) species during the angle spinning (MAS) solid-state NMR (both 1H–13C and 13C–13C)
MTH reaction (Fig. 1a). That means a direct mechanism may be spectroscopy (Fig. 2c). The strong signals at 52.2 (13C) and 3.59 (1H)
operative, at least in the early stages of the MTH reaction16. Since ppm were assigned to surface adsorbed methanol, whereas the sig-
then, research groups (for example, Kondo, Fan, Copéret, Lercher, nals at 57.7 (13C) and 3.54 (1H) ppm were attributed to the SMS (red
Weckhuysen and Liu) have delivered both experimental and theo- strip in Fig. 3a). Interestingly, a strong cross-peak between these 13C
retical evidence in support of the existence of a direct mechanism signals was also observed, which was due to the close proximity of
during the initial stages of MTH11–18,23–30. In essence, this reaction surface adsorbed methanol and the SMS (blue strip in Fig. 3a and
can be sub-divided into two parts: direct mechanism and the con- schematic illustrated in Fig. 2c). Such close proximity does not only
ventional HCP/dual-cycle mechanism, during the induction and reveal the ongoing reaction between SMS and methanol (through
autocatalytic periods, respectively. the polarization of the C–H bond of the SMS by a neighboring adja-
There are several excellent reviews available in the literature cent oxygen atom), but also showcases the carbene/ylide-character
on this topic that adequately describe the history and develop- of SMS.
ments of direct mechanisms4,6. Although a consensus has currently Apart from the carbene mechanism, multiple analogues of oxo-
been reached regarding the existence of the direct mechanism, its nium and methane–formaldehyde-type direct mechanistic propos-
actual nature/course of action is yet to be fully established. In all als have also been shown in the recent literature. The conventional
direct mechanistic proposals, a surface methoxy species (SMS), methane–formaldehyde mechanism of the MTH reaction was origi-
formed upon adsorption of methanol onto a Brønsted acid site, nally postulated by Hutchings and colleagues33 in 1987. The direct
is indisputably the most experimentally verified intermediate C–C bond coupling reaction between methane and formaldehyde,
(Fig. 2)13,14,18,26. The ability of the SMS to form a C–C bond at higher as a result of disproportionation of methanol (through its hydride
reaction temperatures is a well-established phenomenon during any abstraction by the SMS, as illustrated in Fig. 2d), was hypothesized
zeolite-catalysed hydrocarbon conversion process18. The idea of a to pose an unrealistically high energy barrier22. However, several
carbene-insertion mechanism by the SMS was first introduced by recently postulated proposals are conceptually similar to this mech-
Hunger and colleagues18,26,27, and was mostly based on NMR spec- anism. The methoxymethyl cation mechanism proposed by Fan
troscopy and carbene-trapping experiments. The carbene charac- et al.28,29 was developed from the original proposal of Hutchings and
ter of the SMS was proposed to form through the polarization of colleagues, in order to make it energetically feasible. Their experi-
the C–H bond of the SMS by a neighboring adjacent oxygen (as mentally and theoretically verified proposal involves the formation
depicted in Fig. 2a–c)31,32. The carbene/ylide nature of the SMS was of direct C–C bond via: (i) a methoxymethyl cation (CH3OCH2+
confirmed by a carbene trapping experiment with cyclohexane act- from SMS and DME, that is, instead of direct formation of formal-
ing as a probe molecule at ≥493 K, where methylcyclohexane was dehyde), and (ii) its subsequent direct C–C coupling with another
formed through an insertion reaction of a carbene/ylide into the DME/methanol molecule to form CH3OCH2CH2OR (R =H, CH3)
sp3 C–H bond of cyclohexane (Fig. 2a)18. Later, Kondo and col- (a precursor for olefins, as shown in Fig. 2e)28,29. Similarly, the meth-
leagues11,12 also arrived to the same conclusion with the help of yleneoxy mechanism was recently reported by Liu and co-workers14;
IR spectroscopy: the C–C bond containing hydrocarbon species based on in situ solid-state NMR spectroscopy by measuring
CO
O O O O –CH 4 O O CH3OH/ C
Al Al Al j O CH3 O O HCP
CH2O CH3OH
R' = H,CH 3 C +
H CH3
O O Olefin
Methoxymethyl cation mechanism O O
Al
Al
Koch carbonylation mechanism
Methane–Al/oxonium mechanism H2
CH3 CH2 H H
O O O O H C O
Al CH3OCH3 O O H-abstraction O O CH4 O CH3 O O
Al by Al O O Al
k Al – CH2=CH2
O O neighbouring Al
O O O O CH3OH O O
O O
Fig. 2 | Several proposed direct mechanistic routes during the early stages of the zeolite-catalysed MTH process. a,b, The proposed mechanism of the
methanol conversion reaction between the SMS and cyclohexane (a) and ethylene (b) via the carbene-type intermediate. c, The spectroscopically verified
intermediate consists of SMS and methanol, suggesting a possibility of C–H insertion of methanol by carbene-like SMS species. d,e, Schematic illustration
of the conventional methane–formaldehyde mechanism (d), and its modified analogue the methoxymethyl cation mechanism (e). f–i, The methyleneoxy
mechanism of the MTH reaction, involving either the trimethyloxonium (f) or SMS (g) intermediate. The SMS involving route initiates the formation of
both ethylene (h) and methane/formaldehyde (i) simultaneously. j, Plausible routes for the direct formation of the carbon–carbon bond during early
stages of the MTH reaction through Koch-type carbonylation of SMS. k, The methane–aluminium/oxonium mechanism for the conversion of methanol
over alumina.Credit: panel a adapted from ref. 18, American Chemical Society; panels b and c adapted from refs 11,13, Wiley, respectively; panels d and e
reproduced from refs 29,33, Royal Society of Chemistry; panels f–i reproduced from ref. 14, Wiley; panel j adapted from ref. 13, Wiley; panel k adapted from
ref. 17, American Chemical Society.
13
C-methanol conversion over H-ZSM-5 using a rotor reactor active surface–zeolite catalyst (Fig. 2f,g). Although both methane
(Fig. 2f,g & Fig. 3b). Three major bands at 59.5, 69.0 and 80.0 and formaldehyde are proposed to be involved in the catalytic cycle
ppm were generated almost instantaneously and were assigned to (by assisting the regeneration of SMS), their role towards the forma-
the SMS, surface-adsorbed DME and trimethyl oxonium (TMO), tion of the direct C–C bond was not clarified14.
respectively. The appearance of these three bands was accompanied Although numerous mechanistic reports have identified form-
by formation of ethylene (detected by GC-MS) as well as higher aldehyde as a side-product, its fate/role during the MTH reaction
hydrocarbons (chemical shifts of δ =20–40 ppm). Therefore, the is still a subject of debate. Formation of formaldehyde on solid
direct C–C bond was proposed to be formed as a result of direct acid catalysts from methanol is quite predictable, particularly at
interaction between the reactant (methanol/DME) and zeolite- high temperatures. Interestingly, it could be easily associated to
bound surface intermediates: the SMS and TMO (Fig. 2f,g). The the Koch carbonylation mechanism of the MTH reaction (Fig. 2j),
important C–H bond-activation step could be assisted by frame- another direct mechanistic route proposed independently by the
work oxygen to form active methyleneoxy species (R–O–CH2– groups of Lercher and Weckhuysen13,16 at the same time. Lercher
H···zeolite), which eventually leads to the simultaneous formation and co-workers16 first proposed methyl acetate (CH3CO2CH3,
of both an olefin (Fig. 2h) and methane/formaldehyde (Fig. 2i). The derived via carbonylation of methanol/DME) as the very first
unprecedented downfield response and relatively broader nature of C–C bond-containing intermediate during the MTH reaction over
the SMS at 69.0 ppm under in situ reaction conditions (compared H-ZSM-5. Spectroscopic evidence for this proposal came from the
to δ =59.0 ppm under ex situ reaction conditions) was attributed Weckhuysen group13, whom employed a combination of solid-state
to the strong interaction between the activated DME species and NMR spectroscopy (coupled with operando UV-visible diffuse
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
a b CH3
H2 CH3
C H2
H O C O H2
OH H3CO C
H3C C
H H3 H CH3
CH3OH H H H
Zeolite 69.0
C
O O O O
O O Z 80.0 59.5 Z
Al
13 12 min
C–13C 13
C–1H
(57.7, 52.2)
(52.3, 3.59) 5 min
C(ppm)
13
C(ppm)
55 55
(57.7, 3.54)
13
60 60 0 min
60 55 5 4 3
13 1
180 160 140 120 100 80 60 40 20 0 –20 –40
C (ppm) H (ppm)
δc (ppm)
Kubelka–Munk function
0.00
50,000 40,000 30,000 20,000
212
(22.3,2.2)
360
25 (22.3,180.5) 25 1.0 3 h methanol flow
63 min methanol flow
274
C (ppm)
13
295
C (ppm)
30 30
0.5
13
50 50
0.0
Pre-treated
55 (55.2,3.82) 55 50,000 40,000 30,000 20,000
(55.1,178.5)
Wavenumber (cm–1)
180 55 25 20 4 3 2
13 1
C (ppm) H (ppm)
e f 244 144
0.5
248 53
154
203 55 198
0.5 0.4 189 198
141 + 141
*
0.3 185
0.4
0.2
Absorbance (a.u.)
244
58 58
0.2 245 + 245
155
249 ** 150
56
0.1
0.0
300 400 500 600 700 250 200 150 100 50 0 –50
Wavelength (nm) δ (ppm)
Fig. 3 | The spectroscopic signatures of crucial intermediates during the zeolite catalysed methanol-to-hydrocarbon (MTH) process. a, Solid-state
NMR spectra of methoxy in H-SAPO-34 after the MTH reaction for 30 minutes at 673 K. Zooms from 2D 13C–13C (blue strip) and 13C–1H (red strip) magic
angle spinning (MAS) solid-state NMR spectra with long mixing (150 ms) and cross-polarization (CP) contact time (500 ms), respectively, identifying
surface adduct between SMS and methanol (the arrow in the molecular structure indicates 13C–13C correlation). b, In situ 13C MAS solid-state NMR spectra
measured during MTH reaction over H-ZSM-5 at 573 K for 12 min. c, Solid-state NMR spectra of acetate species in H-SAPO-34 after the MTH reaction
for 30 minutes at 673 K. Zooms from 2D 13C–13C (blue strip) and 13C–1H (red strip) MAS solid-state NMR spectra with long mixing (150 ms) or CP contact
time (500 ms), respectively, indicating surface acetate and methyl acetate resonances. d,e, UV-visible diffuse reflectance spectra of H-ZSM-5 (d) and
H-SAPO-34 (e) being exposed to the MTH reaction under in situ (573K) and operando (673K) reaction conditions, respectively. An absorption band at
295 nm (d) on H-ZSM-5 and 387 nm (e) on H-SAPO-34 were assigned to a methylated cyclopentadienium and hexamethylbenzenium ions, respectively.
f, 13C MAS solid-state NMR spectra of retained organic species in H-SAPO-34 (upper spectrum, 573K, 15 min) and H-SSZ-13 (lower spectrum, 548K, 25
min) after continuous-flow of 13CH3OH for 15 min (asterisks represent spinning side bands, chemical shifts on the molecular structures were theoretically
calculated values).Credit: panels a–c from refs 13,14 and 13, Wiley, respectively; panel d reproduced from ref. 46, American Chemical Society; panel e
reproduced from ref. 13, Wiley; panel f reproduced from ref. 53, Wiley
reflectance spectroscopy (DRS) and online mass spectrometry); it that, for ZSM-5, all olefins except ethylene are produced from the
was shown that the Koch-carbonylation route occurred during the olefinic cycle. Ethylene evolution was linked to the presence of
MTH over H-SAPO-34. In this case, the direct C–C bond contain- lower methylbenzenes, suggesting that ethylene is mostly a prod-
ing zeolite-bound acetate species (Fig. 2j) is a Koch carbonylation uct of the aromatic cycle37. Sun and co-workers38 further contrib-
product of the SMS. Here, either methanol or formaldehyde, in uted to this topic by performing seminal kinetic investigations at
principle, could act as a carbonylating agent under MTH reaction different conversion levels in the presence of aromatic and olefinic
conditions. Next, methyl acetate was formed after methoxylation co-feeds. They concluded that both cycles are active for ethylene
of surface-bound acetate species, which independently could initi- and propylene production, with the aromatic cycle giving similar
ate the formation of HCP species and thus, olefins. In the carbonyl selectivity for both olefins. The olefinic cycle, on the other hand,
region of the NMR spectrum, clear cross-peaks to methyl carbon was far more selective to propylene than ethylene, which implies
atoms were observed at both long and short mixing times (Fig. 3c). that, if the two cycles equally contribute to the product distribution,
The signal at 180.5 ppm shows only one cross-peak — correspond- most ethylene will be formed in the aromatic cycle39. The coexis-
ing to a zeolite–acetate species — at 22.5 ppm (Fig. 3c). Moreover, tence of the two cycles naturally renders them as competing40. In
the signal at 178.5 ppm has a clear cross-peak with a 13C signal at this spirit, one can speculate that a desired hydrocarbon range can
55.1 ppm at longer mixing times only (that is, the methoxy carbon be obtained by either stimulating or suppressing one over the over.
is not directly attached to the carbonyl carbon). This 13C-methoxy Such co-catalytic features of HCP species during the MTH reaction
signal correlates with a H signal at 3.82 ppm and has an additional were later theoretically verified by van Speybroeck and colleagues41,
very weak correlation signal with a methyl at 22.3 ppm (13C) at indirectly reinforcing the concept of hybrid organic-inorganic
longer C–C mixing times (Fig. 3c). This cross-peak pattern is a nature of a working MTH catalyst, as originally proposed by Svelle
signature response from a methyl acetate molecule. Thus, all three and co-workers36.
responsible intermediates of this mechanism (SMS, zeolite-acetate UV-visible DRS is possibly the most utilized spectroscopic tech-
and methyl acetate) were spectroscopically identified by solid- nique for the characterization and identification of zeolite-trapped
state NMR spectroscopy. This Koch carbonylation mechanism organics. The biggest advantage of UV-visible DRS is its ability to
of the MTH reaction has recently been theoretically verified by differentiate between carbocationic HCP species and their neutral
Plessow and Studt30. Interestingly, the simultaneous existence of counterparts. For instance, the absorption band of any arenium
(at least) two different direct C–C bond-forming routes (that is, HCP cation is lower in energy than any electronic transition of its
carbene and Koch carbonylation) were spectroscopically identi- neutral counterpart. Moreover, UV-visible DRS provides insightful
fied (Fig. 2c,j & Fig. 3a,c). information regarding the zeolite framework dependent formation
Another interesting recent report by Copéret et al.17 should be of deactivating species during the course of reaction. In general,
mentioned: the formation of the direct C–C bond from DME alter- multiple bands at around ≤295, 340–360, 385–410, 460–500 and
natively proposed to be catalysed by extra-framework aluminium ≥600 nm are observed during a MTH reaction and were typically
atoms in acidic zeolites (Fig. 2k). Herein, the C–C bond forming attributed to neutral benzene/cyclopentadienyl species; dienylic
step initially involves generation of a transient aluminium–oxonium carbocationic/methylbenzeniums (up to three methyl groups);
species (that is, formaldehydic oxygen coordinated to a Lewis acidic highly methylated areniums (specifically hexamethylbenzenium
aluminium) through hydrogen abstraction from an aluminium– ions, HMB+); trienylic and methylated polyarenium ions, respec-
methoxy species. Next, aluminium–oxonium species react with a tively (Fig. 3d,e)25,26,42–47. The formation and characteristics of
methane molecule to yield the direct C–C bond containing surface– these bands are quite unique depending on the zeolite framework
ethanolic species (that is, a precursor for ethylene). This contribu- topology and acidity; for instance, the specific role of alkyl cyclo-
tion inevitably sparks the controversy about the actual involvement pentadienium ions over an MFI zeolite (for example, H-ZSM-5)
of the zeolitic Lewis acid sites during the direct mechanism of the was identified by Jentoft and Wulfers using in situ UV-visible DRS
MTH reaction15,17. However, such discussion is beyond the scope of (Fig. 3d)46. Similarly, the Weckhuysen group identified the nature of
the current review and demands further research to understand the governing active and deactivating methylated aromatic species dur-
phenomenon at the molecular level. ing the MTH process over CHA zeolites (for example, H-SAPO-34
and H-SSZ-13) employing a combination of operando UV-visible
Dual-cycle concept DRS and online gas chromatography/mass spectrometry13,42,47.
After a rather short induction period assigned to the formation of Using a similar strategy involving operando UV-visible DRS, the
the direct C–C bond, the MTH process continues with the steady- same group very recently monitored the formation of active HCP
state formation of hydrocarbons in the so-called autocatalytic species and the accumulation of coke molecules during both
dual-cycle concept34. This notation is the result of merging the two H-ZSM-5 and Mg-ZSM-5-catalysed MTH48. Such spatiotemporal
mechanistic schemes developed in parallel for ZSM-5 and SAPO- UV-visible spectroscopic approach reveals the formation of a coke
345. For cage-like zeolites (SAPO-34), able to accommodate large front at the beginning of reactor bed, which travels towards the end
aromatics, light olefins are believed to form via an indirect way until full deactivation. Magnesium modification resulted in slower
through HCP species (Fig. 4). They can be visualized as typical progression of the coke front and higher olefin selectivity. However,
ship-in-a-bottle molecules that can be formed in zeolites but can- identification/assignment of zeolite-trapped any organic reaction
not desorb. These species are methylated yielding light olefins such intermediates by UV-visible DRS in combination with theoretical
as ethylene and propylene by elimination reactions, thus restoring calculations is not always straightforward and, to some extent, con-
the initial HCP species. For ZSM-5 and its analogues, formation of fusing. For example, HMB+ and the 1-methylnaphthyl cation dis-
hydrocarbons was rationalized as a result of consecutive methyla- play a similar absorption band at ~390 nm (refs 41,44). Unfortunately,
tion and cracking reactions already in the 1980s by Dessau35. The both are probable intermediates of the MTH reaction, whereas
dual cycle represents a compromise between both mechanisms (ole- HMB+ is an active HCP species and the other is formed during the
finic and aromatic cycles) running in parallel (Fig. 4). Both cycles deactivation period41. This is the reason behind the enormous rise
can be subcategorized further to elementary steps described by six of utilization of solid-state NMR spectroscopy in recent years for
types of reactions: methylation and cracking of olefins, methylation the accurate structural elucidation of trapped organics within zeo-
and dealkylation of aromatics, hydrogen transfer and cyclization, lites (vide supra)13,14,17. The combination of solid-state NMR with
the latter two acting as the bridging step between the two cycles8. UV-visible DRS has made significant advancement by the groups of
Using transient switching experiments 12C/13C Svelle et al.36 showed Hunger18,25,43 and Haw49–51.
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
a Aromatics + Paraffins b
Aromatization
c
H2O H 2O
nCH3OH
(CH3)n+1
CH3OH CH3OH
nH2O
CH3OH
Higher
alkenes (CH3)n
Alkanes
H 2O
H2O
(CH3)n+2
CH3OH
Fig. 4 | Steady-state mechanism development of the MTH process. a, Dessau and LaPierre described steady state kinetics of MTH reaction over ZSM-5
by a sequence of olefins methylation followed by their cracking or aromatization to give either olefins or aromatics. b, By contrast, Dahl and Kolboe
investigating MTH over SAPO-34 introduced the HCP concept — an active intermediate of aromatic nature — through which light olefins are formed by
dealkylation. c, The dual-cycle postulates that there are two competing cycles running in the zeolite channels governed by olefins and aromatics, both
acting as co-catalysts for MTH and being active HCP species. Panel a adapted from ref. 35, Elsevier; panel b adapted from ref. 19, Springer; panel c adapted
from ref. 5, Wiley.
Due to its capability to provide information at the molecular Based on this mechanistic knowledge, several strategies can be
level, the utilization of solid-state NMR spectroscopy is increas- put forward to drive selectivity of the process in different directions
ing gradually in the field of heterogeneous catalysis13. Using depending on the desired product. The first approach is related to
13
C-enriched methanol not only significantly increases NMR sen- the manipulation of the organic counterpart, that is, changing the
sitivity, but also allows multi-dimensional solid-state NMR correla- concentration of olefin and/or aromatic species. Obviously, the sec-
tion experiments to be performed in order to construct the accurate ond approach deals with the inorganic element and can be achieved
molecular structures along with structural information of dominant by making use of catalyst engineering. Taking propylene and ethyl-
carbenium ions involved in the HCP of the MTH-mechanism6,18,49. ene (and/or aromatics) yields as the measure of cycle dominance,
For instance, the alkylated cyclopentadienium ion was detected as we will now analyse how certain factors selectively propagate one
the primary HCP species over HMB+ within H-ZSM-5 and H-SSZ- cycle over the other.
13 zeolites during MTH reaction, whereas hexa-/hepta-methylben-
zenium cations are widely acknowledged as the governing HCP Towards improved selectivity in MTH
species within SAPO-type molecular sieves (like SAPO-34, DNL-6) When talking about catalyst and process design, the induction
and H-Beta (Fig. 3f)6,13,24,51–54. This observation was quite consis- period was disregarded as a possible tool to manipulate both
tent with the UV-visible DRS reports by the groups of Jentoft and selectivity and lifetime because for a long time it was considered
Weckhuysen13,42,47. These results demonstrate that the formation of as a negligible part of the entire MTH mechanistic picture. Recent
dominant carbenium HCP species is entirely dependent on the zeo- investigations indicated that formaldehyde — formed through dis-
lite’s framework and acidity. proportionation of methanol during direct C–C bond formation —
In a nutshell, the scientific community now accepts that the causes catalyst deactivation via interaction with aromatic molecules
direct mechanism exists during the early stages of MTO reaction resulting in the formation of polycondensed aromatics55. Besides,
and that HCP species vary depending on zeolite framework, acid- formaldehyde formation is accompanied with the production of
ity and reaction conditions. Numerous recent mechanistic reports CH4 (ref. 56). It thus becomes clear that in order to achieve better
provide solid experimental and spectroscopic evidence and the catalyst stability and selectivity, the induction period should be
simultaneous existence of multiple direct C–C bond forming routes re-engineered to avoid formaldehyde formation. This issue can be
during the MTO reaction is quite likely6. addressed at both the catalyst and reactor level. At the reactor level,
utilization of a continuously stirred tank reactor (CSTR) instead of (Fig. 5a,b)70,71. Window openings are limited by eight atoms: big
a plug flow reactor (PFR) leads to lower local methanol concentra- enough to let short-chain olefins pass through, but sufficiently
tions and lower oxygen containing species, thus mitigating catalyst small as to retain bigger molecules inside of their cages, thus creat-
deactivation57. Moreover, utilization of DME instead of metha- ing an ideal playground to incubate the aromatic-based HCP7. In
nol avoids the potential formation of formaldehyde and results in such an architecture, cavity dimensions can influence the nature
higher methylation rates in comparison to methanol, substantially of the aromatic intermediates (transition-state selectivity)72,
prolonging catalyst lifetime56. At the catalyst level, it was proposed whereas window dimensions impose restrictions for the mol-
that addition of rare-earth oxides, such as Y2O3, should selectively ecules formed inside (product selectivity). Thus, slight differences
decompose formaldehyde thus preventing its further interaction in both dimensions can affect product distribution and propylene/
with aromatic species58. Utilization of DME (DME/water) instead of ethylene ratio, however, such dimensions guarantee high selectiv-
methanol seems to be the best solution, since apart from mitigating ity to short-chain olefins (up to 90%)73. Note that the HCP trapped
catalyst deactivation it decreases the overall heat release (the reac- inside also serves as a scaffold for the formation of polycondensed
tion enthalpy of DME dehydration is lower than that of methanol) species, eventually leading to catalyst deactivation. The challenge
making it more attractive from an industrial perspective59. for this type of zeolites is to achieve steady-state performance
of the aromatic cycle by slowing down coke formation, accom-
Taking advantage of autocatalysis plished by utilization of silicoaluminophosphates (SAPOs) with
Both olefins and aromatics are recognized as competitive co-cata- milder acidity74. On the other hand, steady-state performance
lysts, the excessive presence of a certain product promotes the cycle with a constant coke content can be achieved by utilizing fluidized
from which it originates38,40. So the most straightforward strategy to bed reactors, which additionally provide the advantage of a better
enhance the yield of olefins (propylene and butenes) or aromatics is heat dissipation7. Such reactor design, however, requires applica-
to increase their concentration by co-processing them with metha- tion of fluidizable catalysts; that is, with specific particle size and
nol. Several works report the effect of co-feeding a wide range of density. CHA — the most studied topology — is composed from
olefins and/or aromatics with methanol on the product distribution, cylinder-like cavities with big dimensions able to host aromatic
however, with the main aim to shed light on mechanism details60,61. molecules up to pyrene (Fig. 5a,b)75. Such large dimensions ensure
In a wide range of temperatures and at different methanol con- long lifetime in comparison to other 8MR zeolites. Linked to the
version levels, co-feeding of toluene results in an increase of ethyl- product distribution, the following conclusion can be drawn on
ene and methylbenzene concentration at the expense of propylene the effect of cage dimensions: the more spacious is the cage the
and higher olefins as a result of the aromatic cycle propagation40,62–64. higher is selectivity to propylene72,76,77. Compared at the same
This strategy can be used to respond to the constantly fluctuating temperature (350 and 400 °C), LEV with the smallest dimensions
market demands to produce a mixture of olefins with pre-assigned (7.5 ×6.5 Å) yields higher amount of ethylene, whereas more spa-
C2=/C3= ratio. Olefins co-feeding turns out to be less straightforward cious CHA (10.9 ×6.7 Å) and AEI (12.7 ×11.6 Å) yield propylene76.
and the propagation of the corresponding cycle depends on other The product distribution can be further linked to the confined aro-
parameters. The JGC corporation reported significantly enhanced matics; that is, steric limitations imposed by LEV favour formation
propylene yield (up to 60–70%) when C4–C5 products were recycled of methylbenzenes with limited amount of methyl groups which
with methanol feed at temperatures higher than 843 K (refs 65,66). are responsible for ethylene formation. The pear-shape of AEI is
Co-feeding a small amount of propylene at low conversion levels able to hold bulky aromatics, resulting in an unusually high selec-
at 548–623 K results in a higher contribution of the olefin cycle, tivity to butenes (propylene/ethylene/butane =2.8/1/1.1) at 400 °C
confirmed by higher selectivity towards C3+ products63, whereas at (ref.78). Large cage dimensions do not guarantee high selectivity
723 K and higher conversion levels there is no notable effect of ole- to propylene/butenes if combined with very narrow pores hinder-
fins co-feed40. In the first scenario, methylation reactions are pro- ing diffusion of the latter79. An example is AFX (13.0 ×8.3 Å)
moted over cracking, therefore any inclusion of short chain olefins with very small window-openings (3.4 ×3.6 Å) promoting rapid
such as propylene results in a dramatic increase of C3+ products. At growth of polycondensed species and therefore fast deactivation77.
high methanol conversion levels (>70%), the olefinic cycle becomes Similarly, the more spacious cage of ERI delivers higher selectivity
more important38, whereas further addition of olefins promotes to ethylene in comparison to CHA, which might be due to nar-
both formation of higher olefins, which — with equal success — are rower window openings of ERI (3.6 ×5.1 Å) in comparison to
either cracked or aromatized, thus contributing to both cycles. This CHA (3.8 ×3.8 Å)80.
brings another important conclusion: process parameters such as Medium pore 10-membered ring (10MR) zeolites, on the other
temperature and methanol space–time also contribute to the pro- hand, are composed of straight and/or curved channels (Fig. 5c–f).
motion of one cycle over the other when bed effects come into play. In such structures, the aromatic cycle is only able to operate in
Temperature is another decisive parameter in determining the type the more spacious channel intersections81,82. Choosing zeolites
of prevailing reactions. Higher temperatures promote cracking of with absence of those ensures discarding the aromatic cycle. One-
higher olefins rather than their cyclization, which can be perceived dimensional ZSM-22 with TON topology was the first zeolite for
as promotion of the olefinic cycle. Therefore, to maximize propyl- which suppression of the aromatic cycle was postulated83, the same
ene production, the MTP process is carried out at temperatures observations latter being made for other one-dimensional 10MR
higher than 723 K. zeolites (Fig. 5c,d)84,85. The prevalence of the olefinic cycle com-
bined with product shape selectivity results in the predominant
Influence of zeolite topology on product selectivity formation of C5+ products ranging from 50 up to 75% depending on
The zeolite skeleton dictates its shape-selective properties, allowing the conversion levels. The obtained product mixture rather meets
the participation in the chemical reaction of only those molecules the requirements for gasoline (after hydrogenation step), with the
that are able to fit inside the zeolite pores67. This broad definition can formation of propylene being fairly low. Selectivity to propylene
be further subcategorized into reactant-, product- and transition- can be further improved by optimizing acidic properties of zeolites
state selectivity; thus, being an inherent attribute of zeolitic materi- leading to up to 53% of propylene selectivity at 450 °C (ref. 86), while
als, shape selectivity can be effectively utilized to propagate one cycle optimization of textural properties dramatically prolongs catalyst
over another via product- and transition-state selectivity68,69. lifetime87. For zeolites with intersections, two cycles work in paral-
Small-pore 8-membered ring (8MR) zeolites are composed lel, and propagation of one cycle over the other can be achieved by
of large cavities interconnected by narrow window openings other means (Fig. 5e,f)88.
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
a c e g
b d f h
Fig. 5 | Impact of topology on cycles propagation. a,b, 8-membered ring zeolites are composed of large cavities with small window openings; in such
structures aromatics can be formed but cannot escape the cage. The aromatic cycle is prevailing and such topologies are characterized by high selectivity
to short-chain olefins. c,d, 1D 10-membered ring zeolites cannot accommodate aromatic molecules, therefore the aromatic cycle is suppressed in such
type of topology. The product distribution is characterized by a gasoline range of hydrocarbons. e, The 3D 10-membered ring MFI structure is described
by two types of channels — straight and sinusoidal — forming relatively spacious intersections, which can host the aromatic cycle. f, Both the aromatic
and olefinic cycles are running in zeolite channels and the product is characterized by the presence of a wide range of hydrocarbons. g,h, The channels of
12-membered ring zeolites are rather big and can host aromatic hydrocarbons, therefore the aromatic cycle can run inside the channels of a 1D AFI zeolite.
Logically, further increase of the number of T-atoms in window incorporation of alkaline-earth metals lead to a significant reduc-
openings corresponding to 12-membered ring (12MR) zeolites tion of Brønsted acidity and its effect can be regarded as an increase
allows the aromatic cycle to proceed inside the zeolite channels. in the Si/Al ratio98,99. Yarulina et al.98 have shown that calcium incor-
Therefore, 1D SSZ-24 with tubular channels comparable to the poration results in an almost 10-fold reduction of Brønsted acidity.
size of CHA cavities is almost exclusively selective to aromatics at As a consequence, hydride transfer and cyclization reactions were
low conversion levels (~90%) and >26% at high conversion levels suppressed, resulting in very low amounts of paraffins and aro-
(Fig. 5g,h)89. Besides, wider window dimensions result in the forma- matics and maximizing propylene yield up to 53%. Apart from the
tion of much more alkylated aromatics for zeolite beta and MOR development of the second type of porosity improving diffusional
(mainly penta- and hexamethylbenzene) in comparison to 10MR properties of zeolites, desilication and dealumination of zeolites
ZSM-5 (mainly BTX90). Thus, in 8MR and 12MR zeolites composed can be also viewed as an instrument to dilute acid site density and
of large cavities and channels respectively, the aromatic cycle is pre- inhibit secondary reactions94. Zeolites with similar bulk properties
ferred. Nevertheless, being imprisoned by narrow 8MR windows it can still exhibit substantially different catalytic behaviour, which is
selectively produces short chain olefins in one case, and heavy aro- a side effect of a heterogeneous distribution of aluminium within
matics with no such restrictions. the crystal, also called aluminium zoning. For example, an alumin-
ium-rich rim is characteristic for ZSM-5 crystals synthesized using
Impact of acidity on stability and product selectivity TPA+ as structure-directing agent100. Aluminium-zoning obviously
Concentration, location and strength of acid sites are the three can be considered as a local enhancement of acid density, which, as
most important parameters in defining the overall acidity in zeo- explained above, promotes the aromatic cycle. Hydrothermal syn-
lites. Concentration of Brønsted acid sites is primarily linked to the thesis conditions and the precursor composition mixture can also
amount of Al in the zeolite. A literature survey reveals a linear corre- promote different locations of aluminium within the crystal lat-
lation between propylene selectivity and Si/Al ratio, while a notable tice. Here it should be pointed out that the definition of aluminium
reduction in aromatics selectivity corresponds to the decrease of zoning used in the following describes a gradient in the number
aluminium content in the zeolite (established for MFI but also hold- of aluminium atoms per unit volume, not a distribution gradient
ing for other topologies)91–96. Opposite trends strongly evidence the of aluminium over the available T sites per unit cell. Aluminium
competing nature of the two cycles and that higher acid site density zoning in zeolites has been observed since the 1970s101. Around that
propagates the aromatic cycle (Fig. 6a–d). An increase of alumin- time and in the following years different distributions and effects
ium content enhances the chance of reactant molecules to interact were observed and in 1993 Althoff and co-workers published102 a
with each other, thus increasing the chance to form aromatics and systematic study of the parameters influencing aluminium zoning.
ethylene — the product of the aromatic cycle97. The observed trends Besides influencing desilication and dealumination, because
also account for the improved selectivity to propylene for post-syn- strength and number of catalytically active sites in zeolites affect
thetically modified catalysts with different elements as well demeta- the effective diffusion path of reactants, the presence and dis-
lated zeolites. It has been extensively shown that phosphatation and tribution of aluminium in the zeolite framework also directly
Dou et al. at 450 ºC towards larger coke species. One year later, in 2012, Chen and co-
20 Haw et al. at 400 ºC
workers74 reviewed the effect of acid site density, acid site strength
H H and the role of coke formation on deactivation of SAPO zeolites
and in 2015 Olsbye et al.6 reviewed the MTH process inter alia
H H discussing catalyst deactivation by coke formation.
10
Catalyst deactivation by (hydro-) carbon residues is an impor-
H
tant deactivation pathway in MTH. However, as discussed, after
the formation of the first C–C bond species hydrocarbons are
0 formed according to the autocatalytic dual-cycle concept, which
0 200 400 600 800
implies that hydrocarbon species can be both activating and
SiO2 Al2O3–1
deactivating species6. It is therefore interesting to obtain insights
about the nature and location of the first seeds of coke in zeo-
Fig. 6 | Impact of acidity on cycles propagation. a,b, A literature survey91–96 lites. Intuitively one would expect the first coke species to form
reveals a noticeable increase of propylene (a) and a reduction of aromatics in regions with (i) highest accessibility and (ii) highest activity —
(b) formed along with an increase of the SiO2/Al2O3 ratio for microporous both of which can be influenced by aluminium gradients in the
ZSM-5 zeolites tested in the MTH reaction at 673–773 K and in the catalyst. However, resolving the location of the first coke species
wide range of WHSV. Opposite trends highlight the competing nature formed at the length scale of nanometers and relating it to (local)
of aromatics and olefins. c, A significant reduction of aluminium in the aluminium gradients remains extremely challenging. Recently, the
zeolite framework leads to a decrease of Brønsted acid site (BAS) density, first use of atom probe tomography (APT) to investigate the 3D
consequently propagating the olefinic cycle as a result of a decreased distribution of elements in zeolites at the sub-nm scale has been
amount of secondary reactions leading to olefins cyclization and aromatic reported (Fig. 7)108,109. In 2016 Schmidt et al.109 used APT to study
formation. d, An increase of acid site density results in a higher probability coke formation and its relation to inhomogeneous aluminium dis-
for olefins to interact forming aromatic molecules. A reduction of Brønsted tribution in large ZSM-5 crystals after MTH (using 13C-labelled
acidity can be achieved by post-synthetic modification with alkaline earth methanol to distinguish the formed coke species from contamina-
metals or demetallation procedures (c,d) leading to similar trends. tions). APT samples were taken from the (aluminium rich) surface
and the (aluminium poor) core of the ZSM-5 crystal showing a
affects catalytic performance (Fig. 6). It was shown that several clear correlation of coke formation and Brønsted acid site density.
acid-catalysed reactions preferentially occur at the edges of large The authors reported the presence of carbon clusters (using the
ZSM-5 crystals: in 2007, Roeffaers and colleagues103 applied in situ term cluster as synonym for a group of closely positioned atoms),
fluorescence microscopy to monitor catalytic, condensed-phase which revealed insights into the coke formation mechanism,
reactions in individual zeolite crystals. Specifically, they utilized showing that the coke clusters form preferentially around areas
the (acid-catalysed) formation of chromophores during self-con- with elevated aluminium content, even in the aluminium-poor
densation of furfuryl alcohol to visualize the time-dependent dis- core of the crystal. The median size of the observed 13C clusters
tribution of the fluorescent reaction products. With their study, was around 36–69 carbon atoms, which suggests that these clus-
they showed that the presence of sub-units in large, coffin-shaped ters could contain several occluded aromatic species. Independent
ZSM-5 crystals is relevant to catalysis, and proposed that the of the location of the APT (that is, taken from the surface or core
interfaces between the components act as diffusion barriers. The of the large ZSM crystal), 13C clusters were found in each sample,
authors also observed unusual behavior of certain crystals, show- suggesting that in the beginning of the reaction methanol is pres-
ing intense fluorescence emission confined to the outer surface ent throughout the crystal (Fig. 7). As the MTH reaction and cata-
of specific facets of the ZSM-5 crystals and linked this unusually lyst deactivation progress these large clusters merge into the coke
intense emission to aluminium zoning. In the same year Kox and rich regions observed for example, in the surface of large ZSM
co-workers used in-situ UV-Vis microspectroscopy to study the crystals, that is, in regions of elevated aluminium concentrations.
oligomerization of styrene occurring in the micropores of ZSM-5 These observations were found to be in line with the previously
zeolite crystals104, linking reaction kinetics to the diffusion and suggested mechanism of coke formation in ZSM-5 during MTH
catalytic properties of straight and zigzag channels in large ZSM-5 (see refs6,93–96, among others).
zeolite crystals. One year later Tzoulaki et al.105 studied the dif- Variation in silicon sources was taken as a strategy by Wang
fusion properties of large (aluminium-free) silicalite-1 crystals et al.81 to obtain ZSM-5 zeolites with acid sites located either in inter-
using interference microscopy for monitoring time-dependent sections or in straight and sinusoidal channels. Having the same
concentration profiles during molecular uptake and release and Si/Al ratio, zeolites with acid sites located in channels exhibited
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
a b c
20 μm
3 2
4 20 μm
1
3 2
5 µm
25 μm f
d
e
Time of flight
Local
electrode
Position
sensitive
180 nm
detector
Sample needle
VDC Si Al
13 Laser
C O
150 nm
Fig. 7 | Resolving the location of the first coke species formed during the MTH process using APT. a–f, APT was used to study coke formation and its
relation to the inhomogeneous Al distribution in large ZSM-5 crystals after MTH reaction109. A cross section was prepared from a large ZSM-5 crystal after
MTH reaction using 13C-labelled methanol by focused ion beam (FIB) cutting (a and b). For APT four samples (needles) were taken using FIB cutting from
different regions of the crystal cross-section considering aluminium zoning: high aluminium concentrations are indicated in blue in panel c also showing
the position of the samples investigated by APT. A typical needle is shown in panel d. APT then allowed mapping of the 3D distribution of Si, Al, O, and 13C
atoms within each needle (e). Statistical analysis of the distribution of the 13C atoms revealed the presence of carbon clusters indicating the regions of first
coke formation in the catalyst (f). Figure adapted from ref. 109, Wiley.
comparable to the dimensions of the respective zeolite cages (that is, two cycles — olefinic and aromatic — can be promoted one over
the host molecule). As a result, the lattices of all three zeolites were another and whether one of them can independently exist. The
expanded during the MTH process, as revealed by operando X-ray challenge has been partially addressed: recent catalytic results dem-
diffraction (XRD)113. The expansion of the c-axis of the CHA and onstrate that it is possible to extend catalyst lifetime up to approxi-
LEV lattices (that is, longitudinal direction of their cages) is about mately 1 kg CH3OH per gram zeolite while maintaining high
0.9% and 0.5% during the MTH process, which is more than in the propylene selectivities (>50%). However, it should be further clari-
direction of a- and b-axes (that is, in the direction of the width of fied if deactivating species in these cases are also of aromatic nature.
the cage). On the contrary, relatively further expansion in the direc- Considering that for both cycles the active intermediates act as
tion of a- and b-axes of DDR lattice (that is, 0.5% compared to 0.3% co-catalysts, reaction conditions (temperature and methanol partial
along the c-axis) means it becomes wider during the MTH process pressure), feed composition (that is, the presence of olefinic or aro-
due to coke deposition, whereas both CHA and LEV lattices become matic co-feed), zeolite topology and acidity can be tuned to further
longer. This observation provides further evidence for distinctive enhance performance. Having this said, most of the efforts over the
host–guest chemistry between zeolite and retained organics during last few years have been directed towards the selective formation of
catalysis. To further elucidate the concept of host–guest interactions propylene (the main product from the olefinic cycle). In a potential
during catalysis, the report from Liu et al. should be mentioned72. scenario of lack of ethylene, it would be desirable to apply similar
With the help of isotopic tracing and theoretical calculations, they concepts towards the selective formation of this highly important
showed that the cavity size of different SAPO molecular bearing chemical. How to achieve this objective without compromising
identical 8MR pore openings (that is, SAPO-35, SAPO-34, and catalyst lifetime is however still an open question, as ethylene is the
DNL-6) control the molecular size as well as reactivity of the con- main product of the aromatic cycle, its formation would in principle
fined HCP species, which eventually influences the conversion and always be accompanied by a high rate of formation of aromatics.
selectivity of olefins during the MTH reaction72. Although such con- Along the same line, considering potential fluctuations in the ole-
fined/retained/trapped organics within the zeolite framework are fin market, in an ideal scenario there should be a catalyst whose
believed to be rigid in nature, this is not completely true. Although selectivity can easily be tuned by changing the reaction conditions.
elucidating the reaction mechanisms of zeolite-catalysed MTH Here the best solutions seem to go in the direction of co-feeding the
(Fig. 3a,c) and alkylation of aromatics, the same group adapted dif- required products to promote the cycle of interest. This aspect again
ferent solid-state NMR magnetization transfer techniques that were calls for additional efforts at the reactor level.
previously developed for spectral separation of biomolecules on the In summary, in our opinion, MTH is still a fascinating research
basis of their mobility in order to distinguish between mobile (that topic with a number of scientific and engineering challenges to
is, a molecule or group with fast tumbling or rotation) and rigid be addressed in the near future. Surprisingly, the latter (engineer-
(that is, molecule physisorbed in/on zeolite) species (both were ing) challenges have hardly been touched upon in the open lit-
trapped within zeolite)13,114. However, this host–guest feature of erature, while these have been key in the success of the currently
zeolite catalysis has recently been developed and demands further most applied MTH technology, which relies on the fast deactivat-
research activities to confirm its implications in the actual process. ing SAPO-34 and on fluidized bed technology. On a more funda-
Nevertheless, it contributes to the fundamental understandings of mental note, we are looking forward to seeing the implications that
the zeolite-catalysed hydrocarbon conversion chemistry. the recent discoveries in MTH chemistry may have on other high
temperature hydrocarbon chemistries, such as catalytic cracking
Summary and outlook and even direct methane activation, as well as bifunctional cataly-
After more than 40 years, the methanol-to-hydrocarbons process sis concepts, in which MTH chemistry may play an instrumental
remains as one of the most popular topics of research within the role. Indeed, the combination of zeolites with other functionalities
zeolite catalysis community. The recent industrial implementation, opens new avenues for the direct conversion of CO2 or syngas to
with several new commercial plants already running, has triggered olefins, hydrocarbons and aromatics116,117. In such a bifunctional
even more attention from the academic and industrial communi- systems, CO2 or CO are first hydrogenated to methanol which is
ties. In this line, the interest towards the formation of the direct C–C subsequently converted to hydrocarbons, the zeolite of choice being
bond, which until very recently was believed to be due to impurities, responsible for the second step. This is a rather promising concept,
has risen over the last few years. The combined effort by several yet there are a number of obstacles to overcome, such as overhydro-
research groups helped identify several feasible low-barrier direct genation of olefins to paraffins, high selectivity to CO and methane
coupling pathways for this key mechanistic step, although feedstock associated with metal functionality and how to find a proper cata-
impurities (in principle) could still shorten the induction period115. lyst matching in terms of deactivation.
Among them, formaldehyde formed through disproportionation
of methanol was identified as a deactivating species due to its high Received: 6 February 2018; Accepted: 18 April 2018;
reactivity with aromatics and the undesired formation of polyaro- Published online: 12 June 2018
matic species. This important discovery calls for new reactor con-
cepts where reactant feeding should be performed carefully to avoid
high concentrations of methanol and the consequences thereof,
References
1. Chang, C. D. & Lang, W. H. Process for manufacturing olefins. US patent
including heat effects. US4025576 A (1977).
The next mechanistic step, the dual-cycle, is now better under- 2. Vogt, E. T. C. & Weckhuysen, B. M. Fluid catalytic cracking: recent
stood and can be viewed as the main machine for controlling life- developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 44,
time and selectivity to the desired products. The combination of 7342–7370 (2015).
3. Mitchell, S. et al. Structural analysis of hierarchically organized zeolites.
NMR and UV-visible DRS helped pinpoint not only active inter- Nat. Comm 6, 8633–8647 (2015).
mediates, but also deactivating ones, although there is a fine line 4. Stöcker, M. Methanol-to-hydrocarbons: catalytic materials and their
between them. It has been extensively shown that deactivating spe- behavior. Microporous Mesoporous Mater. 29, 3–48 (1999).
cies can become reaction intermediates if, for example, higher reac- A review covering the main achievements in catalyst design and
tion temperatures are used. Moreover, their reactivity depends on mechanism understanding of methanol-to-hydrocarbons process over
the last century.
zeolite topology: channel/cage dimensions can cause spatial restric- 5. Olsbye, U. et al. Conversion of methanol to hydrocarbons: how zeolite
tions preventing certain intermediates to participate in the mecha- cavity and pore size controls product selectivity. Angew. Chem. Int. Ed. 51,
nistic cycle. In 2012, Olsbye et al.5 posed the question whether the 5810–5831 (2012).
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
A methanol-to-hydrocarbons review dedicated to mechanism 32. Salehirad, F. & Anderson, M. W. Solid-state 13C MAS NMR study of
understanding and describing main parameters affecting selectivty to methanol-to-hydrocarbon chemistry over H-SAPO-34. J. Catal. 314,
hydrocarbons. 301–314 (1996).
6. Olsbye, U. et al. The formation and degradation of active species during 33. Hutchings, G. J., Gottschalk, F., Hall, M. V. Ml & Hunter, R. Hydrocarbon
methanol conversion over protonated zeotype catalysts. Chem. Soc. Rev. 44, formation from methylating agents over the zeolite catalyst ZSM-5.
7155–7176 (2015). Comments on the mechanism of carbon–carbon bond and methane
7. Tian, P., Wei, Y., Ye, M. & Liu, Z. Methanol to olefins (MTO): from formation. J. Chem. Soc. Faraday Trans. 83, 571–583 (1987).
fundamentals to commercialization. ACS Catal 5, 1922–1938 (2015). 34. Plessow, P. N. & Studt, F. Theoretical insights into the effect of the
8. Ilias, S. & Bhan, A. Mechanism of the catalytic conversion of methanol to framework on the initiation mechanism of the MTO process. Catal. Lett.
hydrocarbons. ACS Catal 3, 18–31 (2013). 148, 1246–1253 (2018).
9. Van Speybroeck, V. et al. First principle chemical kinetics in zeolites: 35. Dessau, R. M. & Lapierre, R. B. On the mechanism of methanol conversion
the methanol-to-olefin process as a case study. Chem. Soc. Rev. 43, to hydrocarbons over HZSM-5. J. Catal. 78, 136–141 (1982).
7326–7357 (2014). 36. Svelle, S. et al. Conversion of methanol into hydrocarbons over zeolite
10. Schulz, H. About the mechanism of methanol conversion on zeolites. H-ZSM-5: ethene formation is mechanistically separated from the
Catal. Lett. (2018). formation of higher alkenes. J. Am. Chem. Soc. 128, 14770–14771 (2006).
11. Yamazaki, H. et al. Evidence for a “carbene-like” intermediate during the 37. Bjørgen, M. et al. Conversion of methanol to hydrocarbons over zeolite
reaction of methoxy species with light alkenes on H-ZSM-5. Angew. Chem. H-ZSM-5: on the origin of the olefinic species. J. Catal. 249, 195–207 (2007).
Int. Ed. 50, 1853–1856 (2011). 38. Sun, X. et al. On reaction pathways in the conversion of methanol to
12. Yamazaki, H. et al. Direct production of propene from methoxy species and hydrocarbons on HZSM-5. J. Catal. 317, 185–197 (2014).
dimethyl ether over H-ZSM-5. J. Phys. Chem. C 116, 24091–24097 (2012). 39. Wang, S. et al. Polymethylbenzene or alkene cycle? theoretical study on
13. Chowdhury, A. D. et al. Initial carbon–carbon bond formation during the their contribution to the process of methanol to olefins over H-ZSM-5
early stages of the methanol-to-olefin process proven by zeolite-trapped zeolite. J. Phys. Chem. C 119, 28482–28498 (2015).
acetate and methyl acetate. Angew. Chem. Int. Ed. 55, 15840–15845 (2016). 40. Sun, X. et al. On the impact of co-feeding aromatics and olefins for the
An in-depth mechanistic investigation providing spectroscopic evidence methanol-to-olefins reaction on HZSM-5. J. Catal. 314, 21–31 (2014).
in support of the Koch-carbonylation mechanism of the methanol-to- Seminal kinetic investigations demonstrating the autocatalytic nature of
hydrocarbon reaction. the mechanism and discussing the effect of the feed composition on the
14. Wu, X. et al. Direct mechanism of the first carbon–carbon bond formation dominant reaction pathways.
in the methanol-to-hydrocarbons process. Angew. Chem. Int. Ed. 56, 41. Van Speybroeck, V. et al. Mechanistic studies on chabazite-type methanol-
9039–9043 (2017). to-olefin catalysts: insights from time-resolved UV/Vis microspectroscopy
15. Lercher, J. A. New Lewis acid catalyzed pathway to carbon–carbon bonds combined with theoretical simulations. ChemCatChem 5, 173–184 (2013).
from methanol. ACS Cent. Sci. 1, 350–351 (2015). 42. Borodina, E. et al. Influence of the reaction temperature on the nature of
16. Liu, Y. et al. Formation mechanism of the first carbon–carbon bond and the the active and deactivating species during methanol to olefins conversion
first olefin in the methanol conversion into hydrocarbons. Angew. Chem. over H-SSZ-13. ACS Catal 5, 992–1003 (2015).
Int. Ed. 55, 5723–5726 (2016). 43. Dai, W. et al. Intermediates and dominating reaction mechanism during the
17. Comas-Vives, A., Valla, M., Copéret, C. & Sautet, P. Cooperativity between early period of the methanol-to-olefin conversion on SAPO-41. J. Phys.
Al sites promotes hydrogen transfer and carbon — carbon bond formation Chem. C 119, 2637–2645 (2015).
upon dimethyl ether activation on alumina. ACS Cent. Sci. 1, 313–319 (2015). 44. Hemelsoet, K. et al. Identification of intermediates in zeolite-catalyzed
18. Wang, W. & Hunger, M. Reactivity of surface alkoxy species on acidic reactions by in situ UV/Vis microspectroscopy and a complementary set of
zeolite catalysts. Acc. Chem. Res. 41, 895–904 (2008). molecular simulations. Chem. Eur. J 19, 16595–16606 (2013).
An in-depth review providing versatile reactivity aspects of surface- 45. Qian, Q. et al. Single-particle spectroscopy of alcohol-to-olefins over
methoxy species during zeolite catalyzed hydrocarbon conversion. SAPO-34 at different reaction stages: crystal accessibility and hydrocarbons
19. Dahl, I. M. & Kolboe, S. On the reaction mechanism for propene formation reactivity. ChemCatChem 6, 772–783 (2014).
in the MTO reaction over SAPO-34. Catalysis Lett. 20, 329–336 (1993). 46. Wulfers, M. J. & Jentoft, F. C. The role of cyclopentadienium ions in
This article illustrates the concept of hydrocarbon pool species during methanol-to-hydrocarbons chemistry. ACS Catal 4, 3521–3532 (2014).
zeolite catalyzed methanol-to-hydrocarbon process. 47. Borodina, E. et al. Influence of the reaction temperature on the nature of
20. Song, W., Marcus, D. M., Fu, H., Ehresmann, J. O. & Haw, J. F. An the active and deactivating species during methanol-to-olefins conversion
oft-studied reaction that may never have been: direct catalytic conversion of over H-SAPO-34. ACS Catal 7, 5268–5281 (2017).
methanol or dimethyl ether to hydrocarbons on the solid acids HZSM-5 or 48. Goetze, J. & Weckhuysen, B. M. Spatiotemporal coke formation over zeolite
HSAPO-34. J. Am. Chem. Soc. 124, 3844–3845 (2002). ZSM-5 during the methanol-to-olefins process as studied with operando
21. Lesthaeghe, D., Van Speybroeck, V., Marin, G. B. & Waroquier, M. What UV-vis spectroscopy: a comparison between H-ZSM-5 and Mg-ZSM-5.
role do oxonium ions and oxonium ylides play in the ZSM-5 catalysed Catal. Sci. Technol 8, 1632–1644 (2018).
methanol-to-olefin process? Chem. Phys. Lett. 417, 309–315 (2006). 49. Haw, J. F., Song, W., Marcus, D. M. & Nicholas, J. B. The mechanism of
22. Lesthaeghe, D., Van Speybroeck, V., Marin, G. B. & Waroquier, M. The rise methanol to hydrocarbon catalysis. Acc. Chem. Res. 36, 317–326 (2003).
and fall of direct mechanisms in methanol-to-olefin catalysis: an overview 50. Xu, T. et al. Synthesis of a benzenium ion in a zeolite with use of a catalytic
of theoretical contributions. Ind. Eng. Chem. Res. 46, 8832–8838 (2007). flow reactor. J. Am. Chem. Soc. 120, 4025–4026 (1998).
23. Jiang, Y. et al. Effect of organic impurities on the hydrocarbon formation 51. Haw, J. F. et al. Roles for cyclopentenyl cations in the synthesis of
via the decomposition of surface methoxy groups on acidic zeolite catalysts. hydrocarbons from methanol on zeolite catalyst HZSM-5. J. Am. Chem. Soc.
J. Catal. 238, 21–27 (2006). 122, 4763–4775 (2000).
24. Dai, W. et al. Understanding the early stages of the methanol-to-olefin 52. Li, J. et al. Observation of heptamethylbenzenium cation over SAPO-type
Conversion on H — SAPO-34. ACS Catal 5, 317–326 (2014). molecular sieve DNL-6 under real MTO conversion conditions. J. Am.
25. Dai, W., Wu, G., Li, L., Guan, N. & Hunger, M. Mechanisms of the Chem. Soc. 134, 836–839 (2012).
deactivation of SAPO-34 materials with different crystal sizes applied as 53. Xu, S. et al. Direct observation of cyclic carbenium ions and their role in
MTO catalysts. ACS Catal 3, 588–596 (2013). the catalytic cycle of the methanol-to-olefin reaction over chabazite zeolites.
26. Jiang, Y., Hunger, M. & Wang, W. On the reactivity of surface methoxy Angew. Chem. Int. Ed. 52, 11564–11568 (2013).
species in acidic zeolites. J. Am. Chem. Soc. 128, 11679–11692 (2006). 54. Song, W., Nicholas, J. B., Sassi, A. & Haw, J. F. Synthesis of the
27. Wang, W., Buchholz, A., Seiler, M. & Hunger, M. Evidence for an initiation heptamethylbenzenium cation in zeolite: in situ NMR and theory.
of the methanol-to-olefin process by reactive surface methoxy groups on Catal. Lett 81, 49–53 (2002).
acidic zeolite catalysts. J. Am. Chem. Soc. 125, 15260–15267 (2003). 55. Bollini, P. & Bhan, A. Improving HSAPO-34 methanol-to-olefin
28. Li, J. et al. A route to form initial hydrocarbon pool species in methanol turnover capacity by seeding the hydrocarbon pool. ChemPhysChem 19,
conversion to olefins over zeolites. J. Catal. 317, 277–283 (2014). 479–483 (2018).
29. Wei, Z. et al. Methane formation mechanism in the initial methanol-to- 56. Martinez-Espin, J. S. et al. New insights into catalyst deactivation and
olefins process catalyzed by SAPO-34. Catal. Sci. Tech 6, 5526–5533 (2016). product distribution of zeolites in the methanol-to-hydrocarbons (MTH)
30. Plessow, P. N. & Studt, F. Unraveling the mechanism of the initiation reaction with methanol and dimethyl ether feeds. Catal. Sci. Technol 7,
reaction of the methanol to olefins process using ab Initio and DFT 2700–2716 (2017).
calculations. ACS Catal 7, 7987–7994 (2017). 57. Muller, S. et al. Coke formation and deactivation pathways on H-ZSM-5 in
31. Kazansky, V. & Senchenya, I. N. Quantum chemical study of the electronic the conversion of methanol to olefins. J. Catal. 325, 48–59 (2015).
structure and geometry of surface alkoxy groups as probable active 58. Hwang, A. & Bhan, A. Bifunctional strategy coupling Y2O3-catalyzed
intermediates of heterogeneous acidic catalysts: what are the adsorbed alkanal decomposition with methanol-to-olefins catalysis for enhanced
carbenium ions? J. Catal. 119, 108–120 (1989). lifetime. ACS Catal 7, 4417–4422 (2017).
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATure CATAlysIs Review Article
113. Goetze, J., Yarulina, I., Gascon, J., Kapteijn, F. & Weckhuysen, B. M. Competing interests
Revealing lattice expansion of small-pore zeolite catalysts during the The authors declare no competing interests.
methanol-to-olefins process using combined operando X-ray diffraction and
UV–vis spectroscopy. ACS Catal 8, 2060–2070 (2018).
114. Chowdhury, A. D. et al. Electrophilic aromatic substitution over zeolites
generates Wheland-type reaction intermediates. Nat. Catal 1, 23–31 (2018). Additional information
115. Vogt, C., Weckhuysen, B. M. & Ruiz-Martínez, J. Effect of feedstock and Supplementary information is available for this paper at https://doi.org/10.1038/
catalyst impurities on the methanol-to-olefin reaction over H-SAPO-34. s41929-018-0078-5.
ChemCatChem 9, 183–194 (2017). Reprints and permissions information is available at www.nature.com/reprints.
116. Gao, P. et al. Direct conversion of CO2 into liquid fuels with high selectivity
over a bifunctional catalyst. Nat. Chem 9, 1019 (2017). Correspondence should be addressed to B.M.W. or J.G.
117. Jiao, F. et al. Selective conversion of syngas to light olefins. Science 351, Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
1065–1068 (2016). published maps and institutional affiliations.