You are on page 1of 12

Green Chemistry

View Article Online


PAPER View Journal | View Issue

Improving the efficiency of the Diels–Alder


Cite this: Green Chem., 2017, 19, 237
process by using flow chemistry and zeolite
catalysis†
S. Seghers,a L. Protasova,b S. Mullens,b J. W. Thybautc and C. V. Stevens*a
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

The industrial application of the Diels–Alder reaction for the atom-efficient synthesis of (hetero)cyclic
compounds constitutes an important challenge. Safety and purity concerns, related to the instability of
the polymerization prone diene and/or dienophile, limit the scalability of the production capacity of
Diels–Alder products in a batch mode. To tackle these problems, the use of a high-pressure continuous
microreactor process was considered. In order to increase the yields and the selectivity towards the
endo-isomer, commercially available zeolites were used as a heterogeneous catalyst in a microscale
packed bed reactor. As a result, a high conversion (≥95%) and endo-selectivity (89 : 11) were reached for
the reaction of cyclopentadiene and methyl acrylate, using a 1 : 1 stoichiometry. A throughput of 0.87 g h−1
Received 21st August 2016, during at least 7 h was reached, corresponding to a 3.5 times higher catalytic productivity and a 14 times
Accepted 24th October 2016
higher production of Diels–Alder adducts in comparison to the heterogeneous lab-scale batch process.
DOI: 10.1039/c6gc02334g Catalyst deactivation was hardly observed within this time frame. Moreover, complete regeneration of the
www.rsc.org/greenchem zeolite was demonstrated using a straightforward calcination procedure.

1 Introduction
The Diels–Alder (DA) reaction is universally acknowledged for
its rapid, atom-efficient and clean access to molecular com-
plexity on a small scale. This reaction involves a [4 + 2]-cyclo-
addition between a diene and a dienophile, introducing two
new σ-bonds as indicated in red in Scheme 1.1 This enables
one of the most efficient routes towards six-membered com-
pounds. Controlling the stereo-, regio-, and enantioselectivity
is essential for the synthesis of complex natural products and
pharmaceuticals. Typically, a mixture of endo- and exo-isomers
is obtained (Scheme 1). The endo products, however, are Scheme 1 Stereoselectivity of the Diels–Alder reaction.

mostly obtained as the major products due to favorable sec-


ondary orbital interactions.1,2 Despite its wide use at the
research level, problems associated with reactivity, selectivity
required stoichiometric excess of the reagent, the exothermic
and scale-up still have to be overcome for Diels–Alder reac-
behavior of the main and side reactions, poor selectivity,
tions. The core problems are the instability of the reagents, the
extreme reaction times and temperatures, as well as the use of
(toxic) catalysts.3
a
Indeed, the industrial use of one of the most popular trans-
SynBioC Research Group, Department of Sustainable Organic Chemistry and
Technology, Faculty of Bioscience Engineering, Ghent University, Coupure Links 653,
formations for organic chemists is limited due to safety and
9000 Ghent, Belgium. E-mail: chris.stevens@ugent.be purity concerns. Due to competing, rapid and uncontrolled
b
VITO, Vlaamse Instelling voor Technologisch Onderzoek, Boeretang 200, 2400 Mol, polymerization with a potentially explosive character, the
Belgium scale-up of the DA reaction constitutes an important challenge
c
Laboratory for Chemical Technology, Department of Chemical Engineering and
in view of industrial application. Process engineers tend to
Technical Chemistry, Faculty of Engineering and Architecture, Ghent University,
Technologiepark 914, 9052 Ghent, Belgium
consider the DA reaction by definition in the field of non-
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ scalable transformations and, in many cases, refuse chemical
c6gc02334g routes featuring the DA reaction as a key step.3,4 For instance,

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 237
View Article Online

Paper Green Chemistry

Scheme 2 Buprenorphine synthesis from thebaine.8

of the top 200 pharmaceutical products by US retail sales in isomer, as the interaction between the acid sites of the solid cata-
2010, only one marketed drug is produced industrially using lyst and the electron withdrawing group of the dienophile
the DA reaction.3,5,6 Buprenorphine is a potent narcotic and is reduces the LUMO’s energy. This provides a better overlap with
additionally used to treat opioid addiction.7 The synthesis the HOMO of the diene.1,11g,i Moreover, microporous or meso-
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

from thebaine or oripavine includes a DA reaction step with porous structures also exhibit spatial confinement of the
methyl vinyl ketone (MVK) as a dienophile (Scheme 2).8 reagents, even directing the product selectivity in the one or the
Although little is known about DA reactions on the industrial other direction.11e,f,12 The use of zeolites in combination with
scale, Funel and Abele recently published a review containing MRT is much less investigated.15 However, some research groups
a limited number of industrial relevant DA reactions.3 have performed reactions by combining microreactors and acid
In view of the aim of our research group to facilitate chal- catalysis by zeolites: dehydration of methanol to dimethylether,16
lenging reactions in flow,9 microreactor technology (MRT) is benzylation of benzene with benzyl alcohol,17 Knoevenagel con-
believed to be able to overcome the previously described pro- densation and the epoxidation of 1-pentene.18,19
blems. As a consequence of its small dimensions, the reaction In this work, the reaction between cyclopentadiene (CPD),
environment can be controlled meticulously. The desired reac- by far the most popular diene for DA reactions, and methyl
tion time, temperature and, due to efficient mixing properties, acrylate (MA), both reactive and polymerization prone sub-
stoichiometry can exactly be set. A fast, safe and low cost strates, was chosen as the generic reaction to optimize the DA
increase of the production capacity constitutes the key element reaction under continuous flow conditions (Scheme 3). Among
to address the DA reaction in a continuous flow set-up. the commercially available zeolites (e.g. (US)Y, ZSM-5, Beta and
Moreover, the combination with heterogeneous catalysis tends Mordenite), zeolite Y and Beta type catalysts have the largest
to be a perfect match as the immobilization of a solid catalyst pore diameter and surface area.20 The subtype Ultrastable
in a microreactor simplifies both the recuperation of the cata- Y (USY) is obtained after a dedicated dealumination treatment.21
lyst and the work-up of the product.10 In general, the large pore zeolite types USY and Beta are most
DA reactions are known to be catalyzed by Lewis or often used as a solid acid catalyst in organic synthesis.12,22
Brønsted acids. The replacement of homogeneous catalysts, H-USY was selected as a starting point for this analysis,
such as AlCl3, HF or H2SO4, by heterogeneous analogues, is of however, a batch screening of different commercially available
great interest. In addition to concerns about health, safety and zeolites was carried out.
corrosion, the use of homogeneous catalysts is also Being a hot topic in organic chemistry, the DA reaction has
accompanied by high costs for work-up and disposal of waste already been extensively studied considering a wide variety of
streams. Being solid acids, zeolites, often metal-doped or (expensive) ionic liquids and Lewis acid catalysts, focusing on
-exchanged, and mesoporous silica derived catalysts are exten- maximizing the yields and the endo/exo-selectivity.11j,23–25
sively investigated for DA reactions.11,12 Zeolites consist of SiO4 Although some research aims at inversing this selectivity, e.g.
and AlO4 tetrahedra as elementary building blocks in well- through the use of a bulky Lewis acid catalyst,2a,26 in most
ordered 3D structures. Within the cavities, both Brønsted and cases the maximal formation of the typical major endo-adduct
Lewis acid sites are present.13,14 The use of such solid acid cata- is desired. As the exo-isomer is generally obtained after isomer-
lysts enhances the conversion and the selectivity to the endo- isation from an endo/exo-mixture under basic conditions,27

Scheme 3 Target reaction to develop a continuous flow and heterogeneously catalyzed DA process.

238 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017
View Article Online

Green Chemistry Paper

also this research project is directed towards the endo-isomer reaction using a chiral organocatalyst on a silica support. The
and thus the maximization of the inherent selectivity of a (cata- yield varied from 55 to 95% and an endo/exo-selectivity of
lyzed) Diels–Alder reaction. With regard to the reaction of cyclo- 45 : 55 was obtained at room temperature. However, an excess
pentadiene and methyl acrylate, in many publications, yields of 7 equivalents of cyclopentadiene was used and the need for
are up to 90% and higher, albeit obtained on a small scale and long residence times (up to 25 h) leads to low space–time
often requiring an excess of CPD and long reaction times.11j,24,25 yields (0.0024 mmol g−1 min−1).30
Mostly, the batch processes are limited to an endo/exo-selectivity
of about 80 : 20.24,27c Still, some research groups report an excel-
lent selectivity of 90 : 10 and even higher.11j,25 In conclusion, the
2 Experimental
objectives to be achieved can be set at a yield of at least 90%
and a selectivity of 90 : 10 to be able to compete with the 2.1 Materials
majority of the reported batch investigations. All reagents and solvents were purchased from Sigma-Aldrich
The DA cycloaddition has been performed using MRT Belgium, except for the zeolite materials CBV 500 (Si/Al-ratio
before, mainly taking advantage of the excellent heat transfer, 2.6), CBV 720 (Si/Al-ratio 15), CBV 760 (Si/Al-ratio 30), CBV
pressurized conditions and the rapid optimization ability.28 In 3024E (Si/Al-ratio 15), CBV 21A (Si/Al-ratio 10) and CP814E*
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

an attempt to produce aromatics from a renewable feedstock, (Si/Al-ratio 12.5), which were obtained from Zeolyst
Cheng and Huber reported the production of mixtures of aro- International.20 The zeolite powder was pelletized under a 10
matic compounds using gas-phase DA type reactions between ton press and sieved to obtain a 125–200 µm particle fraction.
different furanic species and olefins over a ZSM-5 catalyst in a Dichloromethane was dried by distillation from calcium
continuous flow fixed-bed reactor in the temperature range of hydride, while tetrahydrofuran (THF) and toluene were dis-
450–600 °C.11b In another paper the DA reaction between tilled from sodium and benzophenone prior to use. Ethyl
cyclopentadiene and crotonaldehyde was performed using a acetate and acetonitrile were dried over molecular sieves. Extra
solid Lewis acid in a microreactor set-up. This solid acid con- dry 2-methyl THF (stabilized, AcroSeal®) was purchased from
sisted of aluminium oxide functionalized silica monoliths. ACROS Organics and was used without any further drying step.
Dichloromethane was used as a solvent at a reactor tempera- Cyclopentadiene was obtained by cracking and distillation of
ture of 37 °C. Conversions of 15 to 80% were reached. The opti- dicyclopentadiene before each experiment. The used zeolite
mized contact time was determined to be 9.4 min, resulting in material was characterised using N2-sorption and thermo-
72% conversion, an endo/exo-distribution of 90 : 10 and a pro- gravimetric analysis (Table 1). These results show a signifi-
ductivity of 0.033 mmol g−1 min−1. Although this work shows cantly higher specific surface area (SSA) and pore volume of
promising results, there is still room for improvement in terms H-Beta. The analysed H-USY and H-Beta species contain a
of conversion, productivity and scalability, especially using comparable water content of about 10 to 15 wt%. Additionally,
acrylates in view of their industrial importance.29 More the acid properties of the commercially available zeolite
recently, Chiroli et al. described a continuous flow Diels–Alder powders were measured using NH3-TPD (Table 2).

Table 1 Overview of the default properties of the commercially available zeolites

Zeolite type Product codea Shape SSAb,c (m2 g−1) Pore volumec (cm3 g−1) Pore radiic (Å) Water contentd (wt%)

H-USY CBV 720 Powder 104 0.24 19.5 14


H-USY CBV 720 Pellets 109 0.38 19.2 14
H-Beta CP814E* Powder 177 0.88 18.2 12
H-Beta CP814E* Pellets 155 0.87 18.1 12
a
Zeolyst International. b Specific surface area. c Determined using N2-sorption. d Determined using thermogravimetric analysis.

Table 2 Acid properties of the commercially available zeolites using NH3-TPDa

Zeolite type Product codeb Si/Al Tl,max c (°C) Th,max c (°C) CAS d (10−2 mol kg−1) ΔHd e (kJ mol−1)

H-Y CBV 500 2.6 262 498 179 110


H-USY CBV 720 15 282 482 47 99
H-USY CBV 760 30 236 423 39 69
H-Beta CP814E* 12.5 322 579 39 116
H-ZSM-5 CBV 3024E 15 253 471 78 139
H-Mordenite CBV 21A 10 264 638 96 176
a
Ammonia temperature programmed desorption. b Zeolyst International. c Low- and high-temperature maximum, respectively corresponding to
the weak and strong acid sites. d Acid site concentration. e Heat of desorption of ammonia.

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 239
View Article Online

Paper Green Chemistry

2.2 Microreactor device tures. The space–time (ST) and site-time (SiT) were obtained as
The X-Cube™ flow reactor is commercially available from the ratio of the catalyst bed mass, respectively the molar
ThalesNano and is shown in Fig. 1. The reagent solutions are amount of acid sites (AS), to the corresponding molar reagent
pumped in the stainless steel tubing by means of two high flow rate.
pressure (HPLC) pumps. Each pump can deliver a 0.1–3 mL Catalyst mass
STðkg cats mol1 substrateÞ ¼
min−1 flow rate. In the employed set-up, a single reagent solu- Molar flow rate reagent
tion and the corresponding pump were used. The reagent solu-
Catalyst mass  AS density
tion was placed under a N2-pressure, using a pressurisation SiTðmol ASs mol1 substrateÞ ¼
module of the commercially available Africa flow system Molar flow rate reagent
(Syrris). The tubing was connected to the insulated CatCart®
(catalyst cartridge). In the commercial system, two cartridges
can be used in parallel. These cartridges can be mounted and
dismounted easily and can be heated from 20 to 200 °C in 3 Results and discussion
5 °C intervals, separately. For all experiments, only one 3.1 Batch solvent screening
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

CatCart® was used after a manual fill with the zeolite material
(loading: 370 mg CBV 720; 350 mg CBV 760; 340 mg CP814E*). In order to develop a sustainable production process, a
Finally, the mixture was pumped to the pressure sensor and thoughtful solvent choice has to be made. A key example of a
the back pressure regulator, to which the outlet is connected. green solvent is without doubt water. The Diels–Alder reaction
The reactor pressure can be set per 5 bar, up to a maximum of is an important example of an organic reaction which can be
150 bar. A schematic representation of the X-Cube™ flow accelerated using water as a solvent.31–33 However, taking a
reactor is given in Fig. 2. The reactor was rinsed with (dry) di- look at the application of the Diels–Alder reaction in organic
chloromethane and subsequently with the reagent solution at synthesis hardly shows any reaction that is not performed in
100 °C and 30 bar before start-up, as the catalyst initially either a chlorinated or an aromatic solvent.34,35 A quick data-
showed a high activity towards cyclopentadiene polymeris- base search demonstrates that dichloromethane is by far the
ation. A full experiment description can be found in the ESI.† utmost used solvent for the cycloaddition of CPD and MA.
Toluene and hexane are located on respectively the second and
third place. Additionally, this reaction has been evaluated in a
2.3 Analysis
wide variety of ionic liquids.36 To the best of our knowledge,
The conversion of methyl acrylate to the end product was only four papers have reported the performance of this reac-
determined by integration of the methoxy-signals of MA tion in water.33b,37 By stirring MA and 2 equiv. of CPD at room
(3.76 ppm), the endo-adduct (3.62 ppm) and the exo-adduct temperature in water for 72 h, the end product was obtained in
(3.69 ppm) in the 1H-NMR spectrum of the crude reaction mix- 89% yield and a endo/exo-distribution of 87 : 13.37a
Unfortunately, combining water as a solvent and a zeolite cata-
lyst is not an option, as both reagents and, if any, end products
accumulate inside the catalyst pore system.
Thus, the evaluation of the DA reaction between CPD and
MA (Scheme 3) was initiated in batch in various organic sol-
vents in order to determine the catalyst potential and the
solvent of choice. Thermogravimetric analysis of the zeolites
showed a water content of about 10 to 15 wt% (Table 1). In
order to remove loosely bound water molecules, the zeolite
material was kept for 24 h in an oven at 100 °C prior to its use.
This constitutes a milder zeolite activation compared to calci-
nations or other high temperature and/or vacuum treatments
found in the literature.11a–c,e–j Among the evaluated dry sol-
Fig. 1 Set-up for the continuous flow Diels–Alder process (left) using a
vents, i.e. dichloromethane (DCM), THF, 2-methyl THF, ethyl
zeolite filled CatCart® (right).
acetate and acetonitrile, the highest conversion and endo/exo-
selectivity were obtained in dry DCM (Table 3). Although the
conversion was higher in dry toluene at reflux temperature, di-
chloromethane was chosen for optimization of the reaction in
a continuous flow device. As a 3 times lower reflux temperature
in dichloromethane still results in more than half of the con-
version in toluene and a significantly higher endo-selectivity, it
was envisaged that dichloromethane had more potential to
Fig. 2 Schematic representation of the X-Cube™ flow reactor in the speed up the reaction and to lead to high conversions and
used set-up (M = Mixer). selectivity. In batch, however, even in the presence of a high

240 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017
View Article Online

Green Chemistry Paper

Table 3 Overview of the obtained results of the batch Diels–Alder selectivity. Next, the influence of the zeolite structure was con-
reaction between 0.1 M CPD and 1 equiv. MA at reflux temperature of sidered and other zeolite types with a similar Si/Al-ratio were
the used solvent. An oven-conditioned USY type zeolite was used as a
evaluated. Although these materials possess a comparable
catalyst (CBV 720, Si/Al 15, 780 m2 g−1)
amount of acid sites, their strength and accessibility can be
1
H-NMR influenced by the catalyst architecture. Using H-ZSM-5 and
H-Mordenite catalysts, low conversions were obtained. Finally,
Loading RTa MA EPb endoc the highest conversion was observed using a H-Beta catalyst
Solvent (wt%) (h) (%) (%) (%)
with an analogous Si/Al-ratio, which could be ascribed to the
Dry DCM 0 2 94 6 79 higher specific surface area and pore volume, as well as to the
Dry DCM 0 3.5 90 10 79
Dry DCM 30 2 64 36 92 acid strength of the sites (Tables 1 and 2).39 The high
Dry DCM 30 3.5 57 43 92 maximum temperature corresponding to the strong acid sites
Dry THF 30 2 85 15 73 indicates that, despite the lower amount of acid sites, their
Dry 2-methyl THF 30 2 77 23 71
Dry EtOAc 30 2 87 13 74 strength is higher compared to the CBV 720 material. The
Dry MeCN 30 2 82 18 81 most promising zeolite H-USY and H-Beta species were
Dry toluene 30 2 36 59 73 applied as a catalyst under continuous flow conditions.
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

a
Reaction time. b End product (both endo- and exo-isomers).
c
Percentage of the endo-isomer in the final DA product.
3.3 Dilute continuous flow experiments
In addition to the problematic scalability, only a moderate con-
catalyst loading (30 wt%), a mere 43% conversion was reached version was reached in batch. Therefore, the development of a
after 3.5 h. continuous flow process was considered. The proper use of a
high catalyst loading, a high pressure environment and
efficient radial mixing, as it occurs in an ideal plug flow, inher-
3.2 Batch catalyst screening
ently related to the proposed microreactor device, is expected
In the next phase of this study, different commercially avail- to lead to an improved result. The batch experiments revealed
able zeolite types were evaluated in the batch mode (Table 4). that dichloromethane was the solvent of choice. For continu-
Among the H-(US)Y family, species with a varying Si/Al-ratio ous flow experiments, the pressure drop over the system
were evaluated. In this way, the best suited balance between depends on the solvent viscosity, catalyst, system set-up and
the amount of acid sites and the acid site strength was investi- pump speed. As indicated by the preliminary flow experiments,
gated within a specific zeolite framework.38 CBV 720 (Si/Al 15) the generated pressure drop over the packed catalyst bed
exhibited the highest conversion. An even higher Si/Al-ratio (CatCart®, see Section 2.2) was too high (>100 bar) when
only resulted in a slightly lower conversion and a similar commercial zeolite powder was used. To avoid this, after
a pelletizing, breaking and sieving pre-treatment, a
125–200 µm fraction was used for all flow experiments result-
Table 4 Overview of the obtained results of the batch Diels–Alder ing in a maximum pressure drop below 5 bar. An overview of
reaction between 0.1 M CPD and 1 equiv. MA in dry dichloromethane for the presented flow configuration can be found in Scheme 4. A
2 h at reflux temperature. Oven-conditioned zeolite material was used
starting solution was prepared, containing 0.5 M of both
as a catalyst in a 30 wt% loading
reagents. This was stored at 0 °C and placed under a N2-
1
H-NMR pressure, to ensure a good pump stability and to avoid system
pressure fluctuations. As it has been stated before that Diels–
a
Zeolite type Product code Si/Al MA (%) EPb (%) endoc (%) Alder reactions are known to benefit from pressurized con-
H-Y CBV 500 2.6 82 18 90 ditions, the continuous flow experiments were performed in a
H-USY CBV 720 15 64 36 92 high temperature and high pressure environment (100 °C,
H-USY CBV 760 30 73 27 91
H-ZSM-5 CBV 3024E 15 89 11 86 30 bar).
H-Mordenite CBV 21A 10 95 5 83 Analogous to the preliminary batch experiments, the zeolite
H-Beta CP814E* 12.5 42 58 90 catalyst was conditioned prior to its use, however, in the con-
a
Zeolyst International. b End product (both endo- and exo-isomers). tinuous flow configuration this drying was performed in situ.
c
Percentage of the endo-isomer in the final DA product. The CatCart® was packed with zeolite USY, after physical

Scheme 4 The continuous flow Diels–Alder process using a zeolite USY catalyst bed.

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 241
View Article Online

Paper Green Chemistry

pre-treatment, and subsequently dried by a N2 flush of 1 h over observed starting 9 h after the process start-up (Table 5,
the heated cartridge at 200 °C. A conversion of MA to the end entry 10).
product of 85% was observed, which decreased to 70% after Because of the volatility of the reagents and the solvent, the
6 h (Table 5, entry 1). Additionally, the catalyst was evaluated end products were isolated as a mixture of endo and exo pro-
without any conditioning procedure (Table 5, entry 2). In this ducts in a nearly quantitative yield via mild rotary evaporation.
way, a conversion of 95% was reached, decreasing to 81% after Upon evaporation with mild heating, a product loss, mainly of
6 h. Quite remarkably, the USY zeolite was thus found to be the endo-isomer, was observed. The purity of the obtained
more active towards acid catalysis without any drying step. mixture of DA products was verified using GC(-MS) and was
Innumerable examples can be found in the literature of acid found to be ≥95%. Under the optimal process conditions, the
catalyzed reactions which require a calcination step or a high Diels–Alder products were isolated in a 94% yield, a 89%
temperature and/or a vacuum step to activate these alumino- selectivity towards the endo-isomer and a GC-purity of 97%
silicate catalysts.11a–c,e–j,40 (Table 5, entry 11). A calculated throughput of 0.87 g h−1 over
On adjusting the dienophile to diene ratio, it was found 7 h was reached, corresponding to a space–time yield of
that a small excess of dienophile (1.2 equiv.) resulted in a 0.29 mmol g−1 min−1. This implies a respectively 9 and 121
nearly quantitative conversion and high selectivity to the endo times higher space–time yield in comparison to the published
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

product, as well as a high stability of the process, which were continuous flow procedures of Sachse et al. and Chiroli
the main goals of the present work (Table 5, entry 4). The et al.29,30 Noteworthily, our results were obtained using just a
process was monitored for 7 h. Noteworthily, the contact time 1 : 1 stoichiometric ratio of diene and dienophile and a contact
(CT), which equals the retention time in the catalyst bed, was time of 1.8 min with the catalyst bed.
barely 3.5 min. This corresponds to a space–time of 444 In an attempt to avoid the use of harmful dichloromethane,
kg cat·s mol−1 CPD. Repetition of these process conditions the second and third best solvents, as it was determined using
showed that the conversion remained at least 94% during the a batch solvent screening (Table 3), were also evaluated under
first 7 hours after start-up. In the resulting DA mixture, a the optimized continuous flow conditions (Table 6).
selectivity of 88% endo-isomer was observed. In comparison, Unfortunately, in the case of toluene or acetonitrile as a
when an excess of CPD was used, a fast deactivation of the solvent, both the conversion and the selectivity towards the
catalyst bed was observed (70% after 4 h, see Table 5, entry 3). endo-isomer were significantly lower in comparison to the use
Doubling the flow rate, corresponding to a contact time of of dichloromethane.
1.8 min and a space–time of 222 kg cat·s mol−1 CPD, or
using dichloromethane without pre-drying (Table 5, entries 5 3.4 Neat continuous flow experiments
and 6), both gave rise to a comparably high conversion Finally, an effort was made to eliminate the need for solvent
initially, but a fast drop in conversion afterwards. It indicates and thus the use of undiluted reagents was evaluated. To avoid
that the performance achieved with the higher space–time any uncontrollable reaction in the reagent reservoir, the choice
corresponds to the establishment of thermodynamic was made to handle two separate reagent streams. This con-
equilibrium. figuration entails a slight excess of cyclopentadiene (1.1
Based on the batch screening of catalysts, other promising equiv.). Despite the higher throughput, a lower conversion and
zeolite catalysts were identified for evaluation under continu- selectivity were obtained (Table 7). Moreover, a small fraction
ous flow conditions. A 1 : 1 CPD : MA stoichiometry could be of the cyclopentadiene dimer (about 6%) was observed on
1
applied to reach a high and stable conversion using a H-Beta H-NMR as a side-product.
catalyst or a H-USY species with a doubled Si/Al-ratio, respect- Presumably, the neat DA reaction mainly takes place in the
ively with a space–time of 390 and 420 kg cat·s mol−1 bulk phase and not on the catalyst surface. This was easily veri-
CPD (Table 5, entries 7 and 8). The better performance of the fied by calculating the acid site concentration in the catalyst
CBV 760 catalyst in comparison to CBV 720 in flow is different bed. This was easily verified by calculating the known acidity
from the results obtained in batch (Tables 4 and 5, entries 2 of the catalyst. The acid site concentration of CBV 720 was
and 7). This can be ascribed to the improvements made in the determined using NH3-TPD to be 0.47 mmol H+ per g. Taking
condition procedure of the catalyst. Next, attempts were made into account the internal concentration of MA, the limiting
to increase the flow rate without affecting the high conversion reagent, of 5.5 M, it appears that there is a catalyst loading of
(Table 5, entries 9 and 10). The optimal balance between con- 9 mol% in the CatCart® or a site-time of 9.5 mol AS·s mol−1
version and throughput was found at a contact time of MA. Therefore, knowing that the reaction also proceeds, albeit
1.8 min, corresponding to 195 kg cat·s mol−1 CPD non-selective, without the catalyst, the catalyst loading is too
(Table 5, entry 9). Moreover, in contrast to the H-USY catalyst, low, leading to a non-catalyzed and thus non-selective back-
the H-Beta catalyzed continuous flow process in dichloro- ground reaction. As a result, the high conversion but low endo-
methane without pre-drying did not show a drop in conversion selectivity can be explained by the high temperature and
within a 7 h time frame (Table 5 entries 6 and 11). These reac- pressure conditions. In comparison, under diluted continuous
tion conditions were selected for the optimal Diels–Alder con- flow conditions (Table 5, entry 4), this catalyst loading is
tinuous flow process. Based on the experimental results at a 99 mol%, corresponding to a site-time of 208.7 mol AS·s mol−1
flow rate of 0.3 mL min−1, deactivation is expected to be CPD. These calculations were performed under the assumption

242 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017
View Article Online

Green Chemistry Paper

Table 5 Results of the continuous flow Diels–Alder process using a zeolite catalyst bed (dry dichloromethane, ξa = 0.35 mL, P = 30 bar, T = 100 °C)

1
H-NMR
Entry Zeolite typeb,c Si/Al FR (mL min−1) CT (min) Equiv. MA [CPD] (M) [MA] (M) Timed (h) EP (%) endo (%)
e
1 H-USY 15 0.1 3.5 1 0.5 0.5 1 85 88
6 70 86

2 H-USY 15 0.1 3.5 1 0.5 0.5 1 95 89


1–5.5 88 89
5.5 83 88
5.5–6.5 81 87

3 H-USY 15 0.1 3.5 0.8 0.6 0.5 1 92 89


1–4 81 88
4 70 87

4 H-USY 15 0.1 3.5 1.2 0.5 0.6 1 96 88


2 98 88
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

3 99 88
4 99 88
5 98 88
6 98 88
7 96 88

5 H-USY 15 0.2 1.8 1.2 0.5 0.6 1 95 89


2 95 90
3 95 89
4 95 89
5 80 89
6 68 88
7 62 88

6f H-USY 15 0.1 3.5 1.2 0.5 0.6 1 93 87


2 96 88
3 99 87
4 96 88
5 93 88
6 84 87
7 80 87

7 H-USY 30 0.1 3.5 1 0.5 0.5 1 93 88


2 93 88
3 93 88
4 93 89
5 93 89
6 93 89
7 92 89

8 H-Beta 12.5 0.1 3.5 1 0.5 0.5 1 93 87


2 94 88
3 94 88
4 94 88
5 94 88
6 94 89
7 94 88

9 H-Beta 12.5 0.2 1.8 1 0.5 0.5 1 95 88


2 95 89
3 95 89
4 95 89
5 95 89
6 93 89
7 93 89

10 H-Beta 12.5 0.3 1.2 1 0.5 0.5 0.7 95 88


1 95 89
2 94 88
3 91 89
4 87 89
5 85 89

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 243
View Article Online

Paper Green Chemistry

Table 5 (Contd.)

1
H-NMR
Entry Zeolite typeb,c Si/Al FR (mL min−1) CT (min) Equiv. MA [CPD] (M) [MA] (M) Timed (h) EP (%) endo (%)
f
11 H-Beta 12.5 0.2 1.8 1 0.5 0.5 1 94 89
2 94 89
3 95 89
4 94 89
5 94 89
6 93 89
7 92 89
a
Void fraction in the catalyst bed, estimated value: 40% of total volume. b The catalyst was used without any drying procedure, unless specified
otherwise. c The corresponding zeolite product code, as obtained from Zeolyst International, can be seen in Table 2. d Time since start-up of the
process. e The catalyst was conditioned by a 1 h N2 flush over the cartridge at 200 °C. f Dichloromethane was not dried before use.

Table 6 Results of the continuous flow Diels–Alder process using a H-Beta catalyst bed, without any drying procedure (CP814E*, Si/Al 12.5,
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

ξa = 0.35 mL, P = 30 bar, T = 100 °C)

1
H-NMR
Entry Solvent FR (mL min−1) CT (min) Equiv. MA [CPD] (M) [MA] (M) Timeb (h) EP (%) endo (%)

1 Acetonitrile 0.1 3.5 1 0.5 0.5 1 66 83


2 64 84
3 62 83
4 61 84
5 59 84
6 57 83
7 55 83

2 Toluene 0.1 3.5 1 0.5 0.5 1 84 86


2 87 84
3 86 86
4 85 85
5 83 86
6 81 84
7 79 85
a
Void fraction in the catalyst bed, estimated value: 40% of total volume. b Time since start-up of the process.

Table 7 Conditions and results of the neat continuous flow Diels–


the influence of the pelletizing pre-treatment, comparative
Alder process using a zeolite USY catalyst bed (FR1 = FR2 = 0.1 mL min−1,
1.1 equiv. CPD, ξ = 0.35 mL, CT = 1.8 min, P = 30 bar, T = 100 °C, batch experiments were carried out. For both the H-USY and
CBV 720) the H-Beta catalyst, the powder and the pelletized form
showed a comparable conversion and selectivity, although in
1
H-NMR the case of the H-Beta species a larger difference was observed
Entry Timeb (h) EP (%) endo (%)
in favour of the powder form (Table 8, entries 2, 3, 5 and 6).
a
1 1 85 70 The deposition of cyclopentadiene on the catalyst surface was
3.5 85 69 determined from the changing CPD : MA ratio, which should
2 1 86 70 theoretically equal one at every moment of the experiment.
a
The deviation of this ratio from one can be ascribed to the
The catalyst was conditioned by a 1 h N2 flush over the cartridge at deposition of cyclopentadiene ( polymers) on the catalyst bed.
200 °C. b Time since start-up of the process.
This process leads to a biased view of the conversion of CPD to
the end products from 1H-NMR and is accompanied by a
decreased CPD : MA ratio. In each batch experiment, the depo-
sition of CPD amounted to 17–30% approximately. For com-
that the physical pre-treatment does not affect the amount of
parison, also in flow a continuous deposition of CPD on the
acid sites, nor their accessibility.
catalyst surface was observed (1–6%). In batch, it appears that
the conversion benefits from the drying of the solvent
3.5 Batch confirmation of the observed trends (Table 8, entries 2, 4, 5 and 7). The continuous flow experi-
In order to obtain a full comparison, the batch experiments ment using the H-USY catalyst indicates that without the pre-
were repeated using the non-dried zeolite material, which lead drying of dichloromethane a sudden accelerated decline in
to an increase of the conversion (Table 8). To gain insight into conversion can be observed (Table 5, entry 6). However, this

244 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017
View Article Online

Green Chemistry Paper

Table 8 Overview of the obtained results of the batch Diels–Alder reaction between 0.1 M CPD and 1 equiv. MA at reflux temperature. Non-dried
zeolite was used as a catalyst in a 30 wt% loading. Conversions are reported as observed on integration of 1H-NMR of the crude reaction mixtures
(RT = 3.5 h)

1
H-NMR

Entry Zeolite typea Si/Al Catalyst form Solvent MA (%) EP (%) endo (%) CPD deposition (mmol g−1)

1 No — — Dry DCM 90 10 79 —
2 H-USY 15 Pelletized Dry DCM 47 53 92 13 (26%)
3 H-USY 15 Powder Dry DCM 46 54 92 15 (29%)
4 H-USY 15 Pelletized DCM 56 44 92 15 (30%)
5 H-Beta 12.5 Pelletized Dry DCM 36 64 92 11 (23%)
6 H-Beta 12.5 Powder Dry DCM 30 70 89 11 (21%)
7 H-Beta 12.5 Pelletized DCM 38 62 91 9 (17%)
a
The corresponding zeolite product code, as obtained from Zeolyst International, can be seen in Table 1.
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

was not seen within a 7 h time frame when the H-Beta catalyst Table 9 Overview of the obtained conversions and endo-selectivity for
was used (Table 5, entry 11). the regeneration
In flow, a high loading and a continuous recuperation of 1
H-NMR
the catalyst results in a high catalytic productivity. Assuming Process Timea (h) EP (%) endo (%)
that a batch process is stopped at 70% conversion after 3.5 h
(Table 8, entry 6) as from then on hardly any change in conver- 1 1 92 89
4 70 87
sion was observed, a catalytic productivity of 5.3 g EP per g
H-Beta could be obtained. In contrast, the 7 h flow process, as 2 1 57 85
presented in Table 5, entry 11, has a catalytic productivity of 2 59 85
18.7 g EP per g H-Beta. Moreover, this productivity is a 3 2 94 89
minimal value, as the catalyst deactivation was not monitored 6 94 89
to its boundaries. Evaluating both the lab-scale batch and con- a
Time since start-up of the process.
tinuous processes at 3.5 h, the flow process has a 14 times
higher production of Diels–Alder end products (0.87 g h−1).
conditions was evaluated (Fig. 3, process 2). The obtained con-
3.6 Spent catalyst characterisation and regeneration
version and selectivity, however, were unsatisfying (Table 9,
Although hardly any deactivation of the used zeolite was process 2). Thus, a more severe reactivation procedure was per-
observed within a 7 h time frame under the optimized con- formed. Again a continuous flow rinse of ethanol was per-
ditions, the regeneration advantage of zeolite catalysis was formed after which the zeolite was calcined for 5 h at 500 °C.
considered. Therefore, to demonstrate the reactivation of the In this way, the catalyst was reactivated and a high conversion
catalyst, an excess of cyclopentadiene was used to accelerate and endo-selectivity was observed again under the optimized
the deposition (Fig. 3, process 1). As can be seen from Table 5, continuous flow conditions (Fig. 3 and Table 9, process 3). A
entry 3 and Table 9, process 1, a small excess of CPD leads to a mixture of end products was isolated in a 92% yield, with a
fast drop in conversion. Next, after the loss of conversion and GC-purity of 99%.
selectivity, the challenge remained to regenerate the catalyst’s In addition to our work, sieve analysis was carried out on
activity. In the first attempt, a 4 h high vacuum treatment at the used pelletized zeolite, which revealed attrition of small
150 °C was carried out, after a continuous flow ethanol rinse. zeolite particles upon reaction (Table 10).
Subsequently, a continuous flow process under the optimized

Table 10 Influence of the process on the particle distribution through


sieve analysis

Particle distribution
(%)
Zeolite
type Si/Ala Conditions <125 µm 125–200 µm

H-USY 15 After reaction 16 84


H-USY 15 After regeneration 13 87
and reaction
H-Beta 12.5 After reaction 18 82
Fig. 3 Process conditions of the regeneration experiments of the used
a
zeolite USY. The corresponding product code can be seen in Table 1.

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 245
View Article Online

Paper Green Chemistry

Table 11 Results of the continuous flow Diels–Alder process of cyclo- lyst or expensive ionic liquids. The process was monitored for
pentadiene and alternative dienophiles using a H-Beta catalyst bed 7 h and a calculated throughput of 0.87 g h−1 was reached,
(CP814E*, dichloromethane, [CPD] = 0.5 M, [DP] = 0.5 M, ξa = 0.35 mL,
corresponding to a 3.5 times higher catalytic productivity and
FR = 0.2 mL min−1, RT = 1.8 min, P = 30 bar, T = 100 °C)
a 14 times higher production of Diels–Alder adducts in com-
Timea Conversionb Yieldc endo parison to the small scale batch process. In contrast to various
Dienophile (h) (%) (%) (%) literature examples, the zeolite material was found to be more
1 94 94 (97) 89 active in acid catalysis without any drying step. Afterwards, the
7 92 end products could be isolated in a straightforward manner in
high yields. The use of neat reagent streams was also evalu-
ated, but resulted in a slightly decreased conversion and a sig-
1 100 99 (97) 89 nificant loss of endo-selectivity. The regeneration of the used
7 100
zeolite catalyst was performed using a 5 h calcination step at
500 °C. Finally, the versatility of this process was demonstrated
1 98d 92 (100) 78 as alternative dienophiles were evaluated, leading to high
7 98d yields.
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

1 84e 84 (96) 54
82e
7
Acknowledgements
1 86 60 f

Financial support for this research from the Fund for
7 90
Scientific Research Flanders (FWO Vlaanderen) and VITO
(Vlaamse Instelling voor Technologisch Onderzoek) and BOF
a
Time since start-up of the process. b The conversion was determined (Bijzonder Onderzoeksfonds) of Ghent University is gratefully
by integration of the signals of dienophile and end product(s) in the acknowledged.
1
H-NMR spectrum of the crude reaction mixtures, unless specified
otherwise. c GC-purity is indicated between brackets. d The conversion
was determined by integration of the signals of diene and end
product(s) in the 1H-NMR spectrum of the crude reaction mixtures.
e
T = 200 °C. f After distillation. Notes and references
1 P. Wothers, N. Greeves, S. Warren and J. Clayden,
Organic Chemistry, Oxford University Press, 2001, pp.
3.7 Expansion to other derivatives
907–924.
To demonstrate the versatile performance of this process, 2 (a) M. Hatano, T. Mizuno, A. Izumiseki, R. Usami, T. Asai,
alternative dienophiles were evaluated (Table 11). Isolation was M. Akakura and K. Ishihara, Angew. Chem., Int. Ed., 2011,
performed using mild rotary evaporation. Methyl vinyl ketone 50, 12189–12192; (b) M. K. Kesharwani and B. Ganguly,
(MVK) gave rise to an excellent yield and selectivity towards the Croat. Chem. Acta, 2009, 82, 291–298; (c) J.-H. Zhou,
endo-isomer. Although this selectivity was slightly lower in B. Jiang, F.-F. Meng, Y.-H. Xu and T.-P. Loh, Org. Lett., 2015,
case acrolein (AL) was used, a high yield of Diels–Alder pro- 17, 4432–4435.
ducts was obtained. In order to obtain a high conversion with 3 J. A. Funel and S. Abele, Angew. Chem., Int. Ed., 2013, 52,
acrylonitrile (AN) as a dienophile, the temperature was raised 3822–3863.
to 200 °C. Finally, dimethyl acetylene dicarboxylate (DMAD) 4 V. Casson, T. Snee and G. Maschio, J. Hazard. Mater., 2014,
also lead to an excellent conversion for 7 hours. 270, 45–52.
5 N. A. McGrath, M. Brichacek and J. T. Njardarson, J. Chem.
Educ., 2010, 87, 1348–1349.
4 Conclusions 6 J. T. Njardarson, http://njardarson.lab.arizona.edu/content/
top-pharmaceuticals-poster (accessed 01/02/2016).
An efficient continuous flow process was developed for the 7 N. D. Campbell and A. M. Lovell, in Addiction Reviews, ed.
Diels–Alder reaction of cyclopentadiene and methyl acrylate. G. R. Uhl, Blackwell Science Publ, Oxford, 2012, vol. 1248,
Using zeolite H-Beta catalysis and high pressure conditions, a pp. 124–139.
high conversion (≥92%) and selectivity towards the endo- 8 (a) A. Machara, L. Werner, M. A. Endoma-Arias, D. P. Cox
isomer (89 : 11) were obtained in just 1.8 min contact time in and T. Hudlicky, Adv. Synth. Catal., 2012, 354, 613–626;
the packed catalyst bed. Moreover, a 1 : 1 stoichiometry diene (b) L. Werner, A. Machara, D. R. Adams, D. P. Cox and
to dienophile was applied. These values are similar to some of T. Hudlicky, J. Org. Chem., 2011, 76, 4628–4634.
the best small scale batch procedures and thus the safely scal- 9 (a) D. R. J. Acke and C. V. Stevens, Green Chem., 2007, 9,
able continuous flow process can compete with previously 386–390; (b) A. Cukalovic, J.-C. R. Monbaliu and C. Stevens,
developed batch processes, as these often require an excess of in Synthesis of Heterocycles via Multicomponent Reactions I,
cyclopentadiene, long reaction times, a (toxic) Lewis acid cata- ed. R. V. A. Orru and E. Ruijter, Springer, Berlin

246 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017
View Article Online

Green Chemistry Paper

Heidelberg, 2010, vol. 23, pp. 161–198; (c) T. S. Heugebaert, J. Q. Zhuang, X. C. Liu, X. M. Liu, X. W. Han, X. H. Bao,
C. V. Stevens and C. O. Kappe, ChemSusChem, 2015, 8, F. X. Chang, L. Xu and Z. M. Liu, J. Mol. Catal. A: Chem.,
1648–1651; (d) F. E. A. Van Waes, J. Drabowicz, A. Cukalovic 2003, 194, 153–167.
and C. V. Stevens, Green Chem., 2012, 14, 2776–2779; 22 (a) A. Corma and H. Garcia, Catal. Today, 1997, 38, 257–
(e) F. E. A. Van Waes, S. Seghers, W. Dermaut, B. Cappuyns 308; (b) S. E. Sen, S. M. Smith and K. A. Sullivan,
and C. V. Stevens, J. Flow Chem., 2014, 4, 118–124. Tetrahedron, 1999, 55, 12657–12698.
10 (a) A. Pohar and I. Plazl, Chem. Biochem. Eng. Q., 2009, 23, 23 D. Nakashima and H. Yamamoto, Org. Lett., 2005, 7, 1251–
537–544; (b) P. Watts and C. Wiles, J. Chem. Res., 2012, 1253.
181–193; (c) J. Wegner, S. Ceylan and A. Kirschning, Chem. 24 (a) I. H. Chen, J. N. Young and S. J. Yu, Tetrahedron, 2004,
Commun., 2011, 47, 4583–4592. 60, 11903–11909; (b) J. R. Harjani, R. D. Singer,
11 (a) C. Bernardon, B. Louis, V. Beneteau and P. Pale, M. T. Garcia and P. J. Scammells, Green Chem., 2009, 11,
ChemPlusChem, 2013, 78, 1134–1141; (b) Y.-T. Cheng and 83–90; (c) G. Imperato, E. Eibler, J. Niedermaier and
G. W. Huber, Green Chem., 2012, 14, 3114–3125; B. Konig, Chem. Commun., 2005, 1170–1172; (d) E. Janus,
(c) J. M. Fraile, J. I. Garcia, D. Gracia, J. A. Mayoral, I. Goc-Maciejewska, M. Lozynski and J. Pernak, Tetrahedron
T. Tarnai and F. Figueras, J. Mol. Catal. A: Chem., 1997, 121, Lett., 2006, 47, 4079–4083; (e) I. Lopez, G. Silvero,
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

97–102; (d) S. Ghosh and N. L. Bauld, J. Catal., 1985, 95, M. J. Arevalo, R. Babiano, J. C. Palacios and J. L. Bravo,
300–304; (e) M. V. Gomez, A. Cantin, A. Corma and A. de la Tetrahedron, 2007, 63, 2901–2906; (f ) D. Sarma and
Hoz, J. Mol. Catal. A: Chem., 2005, 240, 16–21; (f ) S. Imachi A. Kumar, Appl. Catal., A, 2008, 335, 1–6; (g) G. H. Tao,
and M. Onaka, Tetrahedron Lett., 2004, 45, 4943–4946; L. He, W. S. Liu, L. Xu, W. Xiong, T. Wang and Y. Kou,
(g) T. Kugita, M. Ezawa, T. Owada, Y. Tomita, S. Namba, Green Chem., 2006, 8, 639–646; (h) X. P. Xu, M. T. Ma,
N. Hashimoto and M. Onaka, Microporous Mesoporous Y. M. Yao, Y. Zhang and Q. Shen, Eur. J. Inorg. Chem., 2005,
Mater., 2001, 44, 531–536; (h) T. Kugita, S. K. Jana, 676–684.
T. Owada, N. Hashimoto, M. Onaka and S. Namba, Appl. 25 (a) K. Erfurt, I. Wandzik, K. Walczak, K. Matuszek and
Catal., A, 2003, 245, 353–362; (i) G. Mashayekhi, A. Chrobok, Green Chem., 2014, 16, 3508–3514;
M. Ghandi, F. Farzaneh, M. Shahidzadeh and (b) V. Escande, T. K. Olszewski and C. Grison, C. R. Chim.,
H. M. Najafi, J. Mol. Catal. A: Chem., 2007, 264, 220–226; 2014, 17, 731–737; (c) A. Kumar and S. S. Pawar, J. Org.
( j) M. Onaka, N. Hashimoto, Y. Kitabata and R. Yamasaki, Chem., 2007, 72, 8111–8114; (d) B. Mathieu, L. de Fays and
Appl. Catal., A, 2003, 241, 307–317; (k) J. K. Park, L. Ghosez, Tetrahedron Lett., 2000, 41, 9561–9564;
S. W. Kim, T. Hyeon and B. M. Kim, Tetrahedron: (e) K. Mori, T. Hara, T. Mizugaki, K. Ebitani and K. Kaneda,
Asymmetry, 2001, 12, 2931–2935; (l) M. Zendehdel, N. F. Far J. Am. Chem. Soc., 2003, 125, 11460–11461; (f ) W. Stefaniak,
and Z. Gaykani, J. Inclusion Phenom. Macrocyclic Chem., E. Janus and E. Milchert, Catal. Lett., 2011, 141, 742–747;
2005, 53, 47–49. (g) Z. Tang, B. Mathieu, B. Tinant, G. Dive and L. Ghosez,
12 D. E. De Vos and P. A. Jacobs, Microporous Mesoporous Tetrahedron, 2007, 63, 8449–8462.
Mater., 2005, 82, 293–304. 26 (a) K. A. Ahrendt, C. J. Borths and D. W. C. MacMillan,
13 J. Weitkamp, Solid State Ionics, 2000, 131, 175–188. J. Am. Chem. Soc., 2000, 122, 4243–4244; (b) A. Barba,
14 A. F. Masters and T. Maschmeyer, Microporous Mesoporous S. Barroso, G. Blay, L. Cardona, M. Melegari and
Mater., 2011, 142, 423–438. J. R. Pedro, Synlett, 2011, 1592–1596; (c) S. Erol and
15 (a) A. Gavriilidis, P. Angeli, E. Cao, K. K. Yeong and I. Dogan, Tetrahedron, 2013, 69, 1337–1344; (d) M. Hatano
Y. S. S. Wan, Chem. Eng. Res. Des., 2002, 80, 3–30; and K. Ishihara, Chem. Commun., 2012, 48, 4273–4283;
(b) Y. S. S. Wan, J. L. H. Chau, A. Gavriilidis and K. L. (e) K. Ishihara and K. Nakano, J. Am. Chem. Soc., 2005, 127,
Yeung, Microporous Mesoporous Mater., 2001, 42, 157–175. 10504–10505; (f ) Z. J. Jia, Q. Zhou, Q. Q. Zhou, P. Q. Chen
16 (a) Y. Liu, S. Podila, D. L. Nguyen, D. Edouard, P. Nguyen, and Y. C. Chen, Angew. Chem., Int. Ed., 2011, 50, 8638–
C. Pham, M. J. Ledoux and P. H. Cuong, Appl. Catal., A, 8641; (g) T. Kano, Y. Tanaka and K. Maruoka, Org. Lett.,
2011, 409, 113–121; (b) Q. Tang, H. Xu, Y. Y. Zheng, 2006, 8, 2687–2689; (h) K. Maruoka, H. Imoto and
J. F. Wang, H. S. Li and J. Zhang, Appl. Catal., A, 2012, 413, H. Yamamoto, J. Am. Chem. Soc., 1994, 116, 12115–12116;
36–42. (i) S. C. Pellegrinet and R. A. Spanevello, Org. Lett., 2000, 2,
17 N. Candu, M. Florea, S. M. Coman and V. I. Parvulescu, 1073–1076; ( j) W. R. Roush and B. B. Brown, J. Org. Chem.,
Appl. Catal., A, 2011, 393, 206–214. 1992, 57, 3380–3387; (k) J. H. Zhou, B. Jiang, F. F. Meng,
18 G. C. Zhang, X. F. Zhang, J. Lv, H. O. Liu, J. S. Qiu and Y. H. Xu and T. P. Loh, Org. Lett., 2015, 17, 4432–
K. L. Yeung, Catal. Today, 2012, 193, 221–225. 4435.
19 Y. S. S. Wan, J. L. H. Chau, A. Gavriilidis and K. L. Yeung, 27 (a) M. Kanao, A. Otake, K. Tsuchiya and K. Ogino,
Chem. Commun., 2002, 878–879. Int. J. Org. Chem., 2012, 2, 26–30; (b) P. Klein, M. Thyes,
20 http://zeolyst.com/our-products.aspx (accessed 20/11/2015). M. Grosse and K. M. Weber, Preparation of biperiden used
21 (a) W. Lutz, Adv. Mater. Sci. Eng., 2014, 2014(1), 1–20; for treating Parkinson’s disease, by reacting phenylmagne-
(b) V. Patzelova, A. Zukal and J. Kloubek, Chem. Pap. – sium compound and bicycloheptenenyl-substituted propa-
Chem. Zvesti, 1985, 39, 361–367; (c) Z. M. Yan, M. A. Ding, none of high exo/endo ratio, DE10124451 (A1), 2002;

This journal is © The Royal Society of Chemistry 2017 Green Chem., 2017, 19, 237–248 | 247
View Article Online

Paper Green Chemistry

(c) A. Okamoto, M. S. Snow and D. R. Arnold, Tetrahedron, 34 K. C. Nicolaou, S. A. Snyder, T. Montagnon and
1986, 42, 6175–6187. G. Vassilikogiannakis, Angew. Chem., Int. Ed., 2002, 41,
28 (a) S. Abele, S. Hock, G. Schmidt, J. A. Funel and R. Marti, 1668–1698.
Org. Process Res. Dev., 2012, 16, 1114–1120; (b) F. Benito- 35 K.-i. Takao, R. Munakata and K.-i. Tadano, Chem. Rev.,
Lopez, A. J. Mettelalj, R. J. M. Egberink, A. H. Velders, 2005, 105, 4779–4807.
R. M. Tiggelaar, H. Gardeniers, D. N. Reinhoudt and 36 https://scifinder.cas.org/ (accessed 08/08/2016).
W. Verboom, Chim. Oggi, 2010, 28, 56–59; (c) M. Rasheed 37 (a) M. J. Diego-Castro and H. C. Hailes, Tetrahedron Lett.,
and T. Wirth, Chim. Oggi, 2011, 29, 53–55. 1998, 39, 2211–2214; (b) D. A. Jaeger and C. E. Tucker,
29 A. Sachse, V. Hulea, A. Finiels, B. Coq, F. Fajula and Tetrahedron Lett., 1989, 30, 1785–1788; (c) A. Manna and
A. Galarneau, J. Catal., 2012, 287, 62–67. A. Kumar, J. Phys. Chem. A, 2013, 117, 2446–2454.
30 (a) V. Chiroli, M. Benaglia, F. Cozzi, A. Puglisi, R. Annunziata 38 M. J. Remy, D. Stanica, G. Poncelet, E. J. P. Feijen,
and G. Celentano, Org. Lett., 2013, 15, 3590–3593; P. J. Grobet, J. A. Martens and P. A. Jacobs, J. Phys. Chem.,
(b) V. Chiroli, M. Benaglia, A. Puglisi, R. Porta, R. P. Jumde 1996, 100, 12440–12447.
and A. Mandoli, Green Chem., 2014, 16, 2798– 39 A. Funez, J. W. Thybaut, G. B. Marin, P. Sanchez,
2806. A. De Lucas and J. L. Valverde, Appl. Catal., A, 2008, 349,
Published on 25 October 2016. Downloaded on 6/8/2020 5:58:12 AM.

31 S. W. Breeden, J. H. Clark, D. J. Macquarrie and 29–39.


J. R. Sherwood, in Green Techniques for Organic Synthesis 40 (a) P. Botella, A. Corma, M. T. Navarro, F. Rey and G. Sastre,
and Medicinal Chemistry, ed. W. Zhang and B. W. Cue, J. Catal., 2003, 217, 406–416; (b) L. Cerveny, K. Mikulcova
Wiley, 2012, pp. 243–262. and J. Cejka, Appl. Catal., A, 2002, 223, 65–72;
32 J. Clayden, N. Greeves and S. Warren, in Organic Chemistry, (c) M. G. Cutrufello, I. Ferino, R. Monaci, E. Rombi,
Oxford University Press, 2012, pp. 877–908. V. Solinas, R. Magnoux and A. Guisnet, Appl. Catal., A,
33 (a) R. Breslow and U. Maitra, Tetrahedron Lett., 1984, 25, 2003, 241, 91–111; (d) X. C. Hu, G. K. Chuah and
1239–1240; (b) R. Breslow, U. Maitra and D. Rideout, S. Jaenicke, Microporous Mesoporous Mater., 2002, 53, 153–
Tetrahedron Lett., 1983, 24, 1901–1904; (c) P. A. Grieco, 161; (e) N. Srinivas, A. P. Singh, A. V. Ramaswamy,
P. Garner and Z.-m. He, Tetrahedron Lett., 1983, 24, 1897– A. Finiels and P. Moreau, Catal. Lett., 2002, 80, 181–186;
1900; (d) D. C. Rideout and R. Breslow, J. Am. Chem. Soc., (f) G. Takeuchi, Y. Shimoura and T. Hara, Appl. Catal., A,
1980, 102, 7816–7817; (e) R. B. Woodward and H. Baer, 1996, 137, 87–91; (g) J. Wang, J.-N. Park, Y.-K. Park and
J. Am. Chem. Soc., 1948, 70, 1161–1166. C. W. Lee, J. Catal., 2003, 220, 265–272.

248 | Green Chem., 2017, 19, 237–248 This journal is © The Royal Society of Chemistry 2017

You might also like