You are on page 1of 10

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Applied Catalysis A: General 387 (2010) 26–34

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Stereo-selective hydrogenation of 3-hexyne over low-loaded palladium


catalysts supported on mesostructured materials
G. Alvez-Manoli a,∗ , T.J. Pinnavaia b , Z. Zhang b , D.K. Lee b , K. Marín-Astorga c , P. Rodriguez c ,
F. Imbert c , P. Reyes d , N. Marín-Astorga a,∗∗
a
Center for Applied Catalysis, Seton Hall University, 400 South Orange Ave., South Orange, NJ 07079, USA
b
Department of Chemistry, Michigan State University, East Lansing, MI 48824-1322, USA
c
Departamento de Química, Facultad de Ciencias, Universidad de Los Andes, Mérida, Estado Mérida, Venezuela
d
Departamento de Fisico-Quimica, Facultad de Química, Universidad de Concepción, Casilla 160-C, Concepcion, Chile

a r t i c l e i n f o a b s t r a c t

Article history: This paper studies the effect of mesostructured materials in the stereo-selective hydrogenation of 3-
Received 4 May 2010 hexyne at 298 K and 40 psig pressure of H2 over Pd-supported catalysts at different substrate:palladium
Received in revised form 28 July 2010 (S:Pd) molar ratios. The catalysts were prepared by impregnation using a toluene solution of Pd(acac)2
Accepted 30 July 2010
to obtain a metal content close to 1 wt.% over SBA-15 with one-dimensional hexagonal structure, MCM-
Available online 8 August 2010
48 silica with cubic structure and three-dimensional pore system and MSU-␥ alumina with a lathlike
particle morphology. All the supports were characterised by nitrogen adsorption–desorption isotherms
Keywords:
at 77 K, TEM, XRD and H2 chemisorption and TEM measurements. The reactions were found to be zero
Mesoporous
Pd
order with respect to 3-hexyne concentration. The starting 3-hexyne produces primarily cis-3-hexene,
MCM-48 which subsequently is either hydrogenated to hexane or isomerized to trans-3-hexene and 2-hexenes
SBA-15 that are found in very small amounts depending on the nature of the support used. Palladium catalysts
MSU-␥ supported on SBA-15 was the most active and selective catalyst.
Stereo-selectivity © 2010 Elsevier B.V. All rights reserved.

1. Introduction Many studies have examined other mesoporous silica materials,


such as the new families of HMS and MSU silicas with wormhole
Mesoporous materials have been the subject of extended framework structures [16–20]. HMS has been reported to exhibit
research in the field of catalysis [1–7]. These species are defined as high adsorption of light hydrocarbons. This property has been asso-
having pore diameters between 2 and 50 nm. They are anticipated ciated with the high affinity of the HMS framework for the preferred
as being useful for the conversion of higher molecular weight sub- adsorption of alkenes or alkanes [21]. A major breakthrough in this
strates. Catalysis based on these materials should have significant area was the synthesis of SBA-15, which became a very popular
benefits for reactions where diffusion limitations play an impor- material because of its larger framework mesopores, ticker pores
tant role. To overcome diffusion resistance, a substantial effort has walls and higher hydrothermal stability in comparison to MCM-41
been expended on the synthesis, characterisation and catalytic per- [22].
formance of new families of mesoporous silica materials such as One of goals of catalysis is to design catalyst systems that pro-
MCM-41, a one-dimensional hexagonal structure [8–11] and MCM- duce only one desired product out of many other possibilities at
48, a cubic structure with two independent three-dimensional pore high metal/substrate ratio. Green chemistry processes based on
systems [12]. MCM-48 also is an attractive solid for use in vari- catalysis aim to eliminate the production of undesirable products.
ous adsorption and catalysis applications [13,14]. MCM-50 has a Since mesoporous silicas possess lager surface area and uniform
layered structure [15], but it lacks structural stability. pore structures they promise to become excellent supports for
many metals. For example, palladium is one of the more impor-
tant metals in the catalyst family. It is the best catalyst for the
hydrogenation of alkynes. Supported Pd catalysts are commer-
∗ Corresponding author. cially attractive for liquid phase organic reactions [4,6,7,23–25]. Pd
∗∗ Corresponding author at: Department of Chemistry and Biochemistry, Center
nanoparticles as catalysts are introduced in mesoporous materials
for Applied Catalysis, Seton Hall University, 400 South Orange Ave., South Orange,
via incipient wetness impregnation, chemical vapour deposition,
NJ 07079, USA. Tel.: +1 973 761 9000; fax: +1 973 275 2496.
E-mail addresses: gabriela.alvez@shu.edu (G. Alvez-Manoli), ion exchange, etc. [26–28]. The major advantages of supported
norman.marin@shu.edu (N. Marín-Astorga). palladium catalysts are ease of separation and good activity

0926-860X/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2010.07.062
Author's personal copy

G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34 27

under mild process condition, such as low pressure and ambient pH value of the mixture was close to 8. The solid gel was aged in
temperature. a closed vessel for 6 h under ambient conditions to form an as-
The stereo-selective hydrogenation of alkynes to give an alkene made mesostructured precursor, labelled MSU-X alumina. The solid
without geometric isomerization is particularly important in many was then subjected to hydrothermal conditions at 373 K for 24 h to
industrial processes in the absence of isomerization, selective form a surfactant-intercaled boehmite mesophase, labelled MSU-
partial hydrogenation of alkyne produces the cis-alkene. Small S/B. MSU-␥ phases with crystalline ␥-Al2 O3 walls were obtained
amounts of the trans-alkene are formed in these reactions, but cat- by calcining the MSU-S/B at 598 K for 3 h and them at 823 K for 4 h
alytic processes do not lead to the production of the trans-alkene [33,34].
as the primary product [29].
The aim of the present work has been to study the influ- 2.5. Catalysts
ence of mesoporous supports, on the activity and selectivity of
Pd-catalyzed hydrogenation of 3-hexyne to cis-3-hexene (Fig. 1) Calcined forms of mesoporous materials MCM-48, SBA-15 and
in toluene. The specific surface areas were evaluated from the MSU-␥ alumina were impregnated with a solution of Pd(acac)2 in
nitrogen adsorption isotherms and palladium dispersions were toluene in slight excess of the amount required to fill the pore vol-
determined through H2 chemisorption and TEM studies. The liquid ume of the supports. The concentration of Pd(acac)2 was sufficient
phase hydrogenation phase of 3-hexyne was performed at 298 K, to achieve a 1 wt.% loading of Pd. The catalysts were dried at 373 K,
40 psig of H2 pressure at different substrate:palladium ratios (S:Pd). calcined at 673 K under air flow for 4 h and labelled as 1% Pd/MCM-
48, 1% Pd/SBA-15 and 1% Pd/MSU-␥. The catalysts were reduced in
2. Experimental situ in hydrogen at 573 K for 1 h prior to use.

2.1. Materials 2.6. Characterisation

3-Hexyne, poly(ethyleneoxide)–poly(propyleneoxide)–poly N2 adsorption–desorption isotherms were obtained with an


(ethyleneoxide) (PEO–PPO–PEO) Pluronic P123 and tetraethy- automatic Micromeritics apparatus Model ASAP 2010 sorptome-
lorthosilicate (TEOS) were obtained from Aldrich and used without ter at −196 ◦ C using a static adsorption procedure. Samples were
further purification. The solvent toluene was obtained from degassed at 150 ◦ C and 10−6 Torr overnight prior to analysis.
Pharmco and was distilled under argon atmosphere before use. BET surface areas were calculated from a linear part of the
Cetyltrimethylammonium bromide (CTAB) and sodium hydroxide BET plot according to IUPAC recommendations. Pore-size distri-
were obtained from MERCK, Ammonium hydroxide (Fisher Sci- butions were calculated from the adsorption isotherms by the
entific), HCl (Riedel de Häen) and hydrogen (99.995% purity, AGL) Barrett–Joyner–Halenda (BJH) model and the total pore volumes
were used as obtained. A commercial catalyst containing 1% Pd were estimated from the N2 uptake at P/P0 = 0.995.
supported on Al2 O3 was obtained from Aldrich. Powder X-ray diffraction patterns were measured using Cu-
K␣ radiation ( = 0.154 nm) and a Rigaku Rotaflex diffractometer
2.2. MCM-48 equipped with a rotating anode operated at 45 kV and 100 mA.
Counts were accumulated every 0.02◦ (2) at a scan speed of 1◦
The synthesis of MCM-48 was performed using a 2/min.
standard procedure [30,31]. Basically, MCM-48 was pre- Metal dispersion was determined by H2 chemisorption at
pared from a gel of the following molar composition 343 K in automatic Micromeritics apparatus Model ASAP 2010
TEOS:Na2 O:CTAB:H2 O = 1:0.25:0.65:0.62. A mixture of deionised chemisorption. Hydrogen chemisorption was carried out at 298 K in
water, cetyltrimethylammonium bromide, tetraethylorthosilicate the pressure range 1–100 mm Hg. The hydrogen uptake was evalu-
and sodium hydroxide was stirred at room temperature for 1 h. ated from the irreversible amount of adsorbed H2 as the difference
The synthesis gel was heated at 383 K for 72 h in an autoclave to between the first (total) and the second (reversible) isotherms in
obtain MCM-48 silica. The final product was filtered, washed with the pressure range 1–4 mm Hg. Transmission electron microscopic
water, dried at room temperature, and calcined at 823 K for 6 h. (TEM) micrographs were obtained in a Jeol Model 100CX micro-
scope equipped with a CeB6 filament and an accelerating voltage
2.3. SBA-15 of 120 kV. The specimen was loaded onto a holey carbon film that
was supported on a copper grid by dipping the grid into a sample
Pluronic P123 (4 g) was added to 30 ml of water, and then 115 ml suspension in ethanol.
of 2 mol L−1 HCl solution was added to dissolve the surfactant. The The Pd contents of the catalysts were evaluated by molecular
surfactant solution was heated to 308 K. Tetraethyl orthosilicate absorption at  = 462.5 nm using a Varian UV–Vis spectrophotome-
(TEOS, 8.5 g) was added slowly to the surfactant solution with stir- ter. Prior determination, the sample were dissolved in aqua regia.
ring at 308 K and the resulting mixture was allowed to age for 20 h. The palladium concentration was obtained by means of a calibra-
The reaction flask was then sealed and aged at 353 K overnight tion curve.
without stirring. The resulting white precipitate was filtered out,
washed with copious amounts of H2 O, and allowed to air dry at 2.7. Catalytic experiments
room temperature for 24 h. The surfactant was removed by calci-
nation in air at 773 K for 4 h [27,32]. The stereo-selective hydrogenation of 3-hexyne in toluene solu-
tion was carried out in a glass reactor at 298 K under different
2.4. MSU- alumina H2 pressures. For each measurement, 25 mg of the powdered cat-
alyst, toluene and 3-hexyne were placed in the reaction reactor.
A solution containing 0.10 mol of the desired aluminium precur- Hydrogenation was conducted under efficient stirring (1500 rpm)
sor was mixed with the desired amount of surfactant in a blender to eliminate diffusion control. The total volume used was 15 ml.
for 5 min. The resulting white mixture was aged in a shaking bath Different substrate:Pd molar ratios were used. Prior to reaction,
at a temperature below 318 K for 36 h to form a clear solution. the catalyst was reduced at 573 K for 1 h under H2 flow and then
Ammonium hydroxide was slowly added under stirring to bring cooled to the reaction temperature. Then, an aliquot of 3-hexyne
the total OH− /Al3+ molar ratio to a value between 3 and 3.6. The in toluene was injected into the reactor under constant stirring.
Author's personal copy

28 G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34

Fig. 1. Reaction scheme for the stereo-selective hydrogenation of 3-hexyne.

Hydrogen consumption was recorded and the reaction products nal mesoscopic organization (p6mm) in agreement with the results
were analysed using a Hewlett Packard 5890 gas chromatograph reported by Zhao et al. [22,32].
equipped with a capillary column DB-1 (0.25 mm; 30 m) and flame The XRD data for the supported Pd catalysts (Fig. 2) show that
ionisation detector (FID). The substrate:palladium ratios (S:Pd) the framework structures of the 1% Pd/SBA-15 and 1% Pd/MCM-
were expressed as mole 3-hexyne:mole Pd. Turnover frequencies 48 silica’s are retained upon metal impregnation, whereas the
(TOF) were normalised according to the number of palladium sites mesostructural integrity in the 1% Pd/MSU-␥ catalyst looks to be
obtained by hydrogen chemisorption analysis. compromised by the process used to impregnate the palladium,
as evidenced the disappearance of the diffraction peak at 2 = 2◦ .
3. Results and discussion The wide angle XDR pattern of the 1% Pd/SBA-15 catalyst shows
a broad peak near 2 = 25◦ corresponding to the Si–O bond length
3.1. Structural and textural properties of the supports of an amorphous framework, along with a weak diffraction peak
near 2 = 34◦ corresponding to the second most intense peak of
Fig. 2 depicts the XRD patterns for calcined MSU-␥ alumina, the metal. The strongest metal reflection is obscured by the sup-
SBA-15 silica and MCM-48 silica mesostructures. Low angle X-ray port. The metal diffraction peaks for the 1% Pd/MCM-48 and 1%
diffraction at (2 from 0.5◦ to 10◦ ) for MSU-␥ alumina shows only Pd/MSU-␥ were more difficult to resolve.
one broad diffraction peak in the small angle region near 2 = 2◦ , TEM images (Fig. 3) verify the regular pore structure of the
which is characteristic of a wormhole or spongelike framework MCM-48 and SBA-15 supports, as well as the disordered channel
[16,33–39]. The MCM-48 cubic phase exhibits four well-resolved arrangement and wormhole structure assignments for the MSU-
hkl diffraction lines consistent with the expected framework struc- ␥ alumina [33]. The TEM micrographs confirm the retention of
ture for siliceous MCM-48 [30,40–42]. The diffraction peaks related the framework structure after Pd incorporation for the 1% Pd/SBA-
to the (2 1 1) and (2 2 0) planes possess higher intensity for syn- 15 and 1% Pd/MCM-48 catalysts in accord with the XRD patterns
thesised sample, which can be taken as an indication of a higher (Fig. 2). However, the XRD pattern shows the presence of the
ordering in the textural uniformity of cubic (Ia3d) structure. The gamma phase for the MSU-␥ alumina. Although the metal impreg-
hexagonal mesoporous SBA-15 support exhibits one intense peak nation process disorders the mesostructure of the alumina support,
at ∼0.95◦ and weaker peaks between 1.4–1.6◦ and 1.7–1.9◦ , which the textural properties for this solid did not show substantial
correspond to 1 0 0, 1 1 0 and 2 0 0 reflections, respectively. The changes (see Table 1) [33].
intensity of the 1 0 0 diffraction line corresponds to a d spacing The textural properties of the mesophases and nitrogen
between 10.2 and 10.4 nm, a value consistent with the reported adsorption–desorption isotherms are provided in Table 1 and Fig. 4
in the literature [2,22,31]. The 1 0 0, 1 1 0 and 2 0 0 Bragg reflections respectively. All the isotherms are type IV, characteristic of well-
confirm that as-synthesised SBA-15 has a high degree of hexago- defined mesoporous frameworks, with a type H1 hysteresis loop
Author's personal copy

G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34 29

Fig. 2. X-ray diffraction patterns of the mesostructured supports and supported Pd catalysts.

[3,16,2,32,43]. The surface areas and specific pore volume of the region below a partial pressure of 0.30, decreased in the order
SBA-15, MCM-48 and MSU-␥ Pd containing materials showed only MCM-48 > SBA-15 > MSU-␥. Due to these surface area changes in
a slightly decrease with the Pd impregnation. These results suggest the catalysts, it can be concluded that Pd introduction in MCM-
that the loss in surface area and pore volume observed by nitro- 48 and SBA-15 does not significantly affect the original structure
gen adsorption measurements are a consequence of pore filling, in agreement with XRD results, but in the MSU-␥ case the Pd
perhaps through a thickening of the framework walls in according impregnation affect the mesostructure order while retaining the
with TEM images (Fig. 3). textural properties. Note that the catalysts have framework pore
Hicks et al. [44] reported that the lathlike morphology in MSU- sizes are in the range 3–7 nm, large enough to accommodate the
␥ catalyst is retained under HDS reaction conditions, suggesting 3-hexyne and the reaction products. Crocker et al. [45], reports
that the loss of surface area and pore volume are a consequence the critical dimension of alkynes, essentially determined by the
of pore filling with retention of the framework. This evidence is thickness of the benzene ring, is 0.3 nm. It is therefore suggested
supported by the fact that the low angle diffraction line indica- that the apparent substrate size dependence arises from spatial
tive of a regular pore–pore correlation length is lost following requirements for reactant chemisorption both on the external sur-
HDS catalyst. The surface areas, as determined by fitting the BET face sites and in the restricted internal region of the mesoporous
equation to nitrogen adsorption isotherms in the partial pressure particles.

Table 1
Textural properties of catalysts and catalyst supports.

Sample SBET (m2 g−1 ) dp (nm) Vp (cc/g) Pd (%)a

MSU-␥ alumina 319 10.1 0.81 –


1% Pd/MSU-␥ 286 7.1 0.76 0.98
SBA-15 860 4.6 1.20 –
1% Pd/SBA-15 793 7.0 1.09 0.99
MCM-48 1184 3.2 0.95 –
1% Pd/MCM-48 1085 2.1 0.90 0.98
a
Determined by molecular absorption.
Author's personal copy

30 G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34

Fig. 3. Transmission electron micrographs of (a) 1% Pd/SBA-15, (b) 1% Pd/MCM-48 and (c) 1% Pd/MSU-␥ catalysts.

3.2. Characterisation of Pd catalysts these reasons the 1% Pd/SBA-15 catalyst provides more uniform
particles of palladium (see Fig. 3a).
The size distributions of palladium crystallites for the palla- The relative amounts of active surface Pd sites on the
dium catalysts have been investigated by TEM measurements. Fig. 3 mesoporous materials were calculated from H2 chemisorption
shows the TEM micrographs for the Pd on SBA-15, MCM-48 and experiments for the Pd-supported catalysts used in the semi-
MSU-␥ catalysts revealed a good narrow palladium particle size dis- hydrogenation of 3-hexyne. The palladium dispersions, obtained
tribution for all catalysts. The size of the palladium particles formed from H2 chemisorption measurements, are displayed in Table 2. It
were in the range 7–10, 1.8–2.7 and 3–5 nm and the average crystal- may be observed from the Table 2 that the dispersion obtained for
lite size was 7.6, 2.4 and 3.8 nm for 1% Pd/SBA-15, 1% Pd/MCM-48 1% Pd/SBA-15 catalyst is lower than the dispersion found for 1%
and 1% Pd/MSU-␥ catalysts respectively. No Pd aggregates larger Pd/MCM-48 and 1% Pd/MSU-␥ catalysts. This behaviour suggests
than 9 nm were detected for any catalyst. the presence of residues on the palladium sites formed during cal-
On the other hand, the palladium particles were confined into cination of the surfactant and acac complex. The carbon coating
the hexagonal channel structure for 1% Pd/SBA-15 and 1% Pd/MCM- on the palladium particles would result in a significant decrease of
48 catalysts (Fig. 3a and b). The uniform dispersion of Pd particles the ability to chemisorb hydrogen. The confinement of Pd inside the
is due to the homogeneous distribution of palladium salt during channel may provide more resistance to sintering during the reduc-
the impregnation process. Palladium on the 1% Pd/MSU-␥ catalyst tion under hydrogen flow. Also, low hydrogen chemisorption by the
(Fig. 3c) is well dispersed and comparable to the pore size of the catalysts at low metal loading (1%) may be due to the formation of
framework (see Table 1), though the metal particles are more dif- Si–O groups covering the palladium particles, resulting in an inhibi-
ficult to resolve in the image due to the disorder of the framework tion of hydrogen chemisorption. On the other hand, the palladium
structure. The Pd is supported on SBA-15 within the framework particle size calculated from H2 chemisorption are in reasonable
pores (see Fig. 3a), as indicated by correspondence between the agreement with the values obtained by TEM (see Table 2).
particle size of the metal and the size of the pore (see Table 1). The
pore occupancy is random in the 1% Pd/SBA-15 catalyst. Pd also is
placed within the pores of 1% Pd/MCM-48 catalyst (Fig. 3b), but the Table 2
H/Pd ratios obtained from chemisorption data and metal particle size of supported
particles are larger than the framework pore size. This may be due
Pd catalysts.
to the particles occupying intersections of the branched pore sys-
tem or to damage of the pore walls upon aggregation of the metal Catalysts H/Pd dH2 (nm) dTEM (nm)
atoms. 1% Pd/SBA-15 catalyst has thicker walls than 1% Pd/MCM-48 1% Pd/MSU-␥ 0.83 1.2 3.8
catalyst and should be less susceptible to damage. Also, the meso- 1% Pd/SBA-15 0.15 6.5 7.6
pores of 1% Pd/SBA-15 catalyst are straight, without branching. For 1% Pd/MCM-48 0.26 3.7 2.4
Author's personal copy

G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34 31

Fig. 5. Effect of catalyst weight on the conversion of 3-hexyne to 3-hexene


semi-hydrogenation over 1% Pd/MCM-48 catalyst in toluene as solvent. T = 298 K,
PH2 = 40 psig, S/Pd = 1000, time of reaction = 20 min.

In an attempt to improve the rate of hydrogenation, the effect


on increasing pressure was studied at room temperature and a sub-
strate:Pd (S/Pd) molar ratio of 1000 (Fig. 6). Increasing the hydrogen
pressure from 10 to 40 psig resulted in an improvement of activity
with slight loss of stereo-selectivity for all Pd-supported catalysts.
It can be seen that the hydrogen compete for the adsorption sites
for the three catalysts studied. This increase in pressure will reduce
mass transfer limitations and allow for improved reactions rates.
We therefore decided that further measurements should be con-
ducted at 40 psig of hydrogen pressure. A similar behaviour was
observed for 2-pentyne hydrogenation by Bennett et al. [46] who
observed a slight decrease in selectivity with increased hydrogen
Fig. 4. Nitrogen isotherms of supported Pd catalysts formed by impregnation with
Pd(acac)2 . pressure.
The stereo-selective hydrogenation of 3-hexyne on Pd-
supported catalysts was carried out at 298 K and 40 psig in toluene
3.3. 3-Hexyne hydrogenation

The presence of a solid catalyst complicates the transport pro-


cess of gaseous and liquid reactant. That is, liquid to solid transport
as well as the transport of components into the catalyst pores
are involved. In the production of fine chemicals, complex multi-
step organic reactions are involved and the selectivity towards the
desired product might be affected by mass transfer limitations.
Mass transfer can play an important role in liquid phase semi-
hydrogenation using mesoporous catalysts. Therefore, it is critically
important that mass transfer effects are ruled out before attempt-
ing to obtain reliable kinetic data. In order to minimise gas-to-liquid
phase mass transfer limitations, the highest stirring speed was used
(1500 rpm). This precaution served to increase the rate of mass
transfer from the gas to the liquid phase, as well as to increase
the rate of substrate transfer from the bulk liquid to the catalyst
surface, and to minimise internal diffusion resistance [29]. To deter-
mine if mass transfer was not rate controlling the conversion of
3-hexyne, the reaction was studied under standardized conditions
over a range of different catalyst masses for all palladium supported
catalyst. The absence of mass transfer limitations at 1500 rpm of
stirring was confirmed by a straight line obtained in the plot con-
Fig. 6. Effect of pressure on the initial rate and 3-hexene product selectivity in the
version against catalyst mass at the same reaction time (Fig. 5). This hydrogenation of 3-hexyne over supported Pd catalysts in toluene suspension. T =
behaviour was exhibited for all studied catalysts. 298 K, S/Pd = 1000.
Author's personal copy

32 G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34

Table 3
cis-Hexene product selectivity and turn over frequency (TOF) for the supported Pd catalysts with different substrate:Pd molar ratios (S:Pd).

Catalyst Selectivity to cis-isomer (%)a TOF (s−1 )


S:Pd molar ratio S:Pd molar ratio

11,000 9000 7000 5000 11,000 9000 7000 5000

1% Pd/SBA-15 98.7 98.9 99.0 99.1 2538 2639 2086 2569


1% Pd/MCM-48 98.2 98.6 98.7 98.6 730 738 854 722
1% Pd/MSU-␥ 98.4 98.4 98.5 98.5 111 135 111 128
a
Conversion level of 50% at T = 298 K, PH2 = 40 psig.

at different S:Pd molar ratios, which tends to affect the activity and
selectivity. The effect of the S:Pd molar ratio on the activity and
selectivity to cis-hexene isomer for the three palladium catalysts
are displayed in Table 3. Table 3 demonstrates that irrespective of
the S:Pd molar ratio, cis-3-hexene formation was predominant for
all Pd catalyst studied. Further increases in the S:Pd molar ratio did
not have a significant effect on the product distribution for all the
catalysts. The turnover frequencies (TOF) displayed no appreciable
variation with the S:Pd molar ratio (see Table 3). It is remarkable
that the 1% Pd/SBA-15 catalyst exhibits a much higher activity at all
S:Pd molar ratio used in comparison with the others Pd-supported
catalyst. The observed behaviour may be attributed to the meso-
porous framework effect, a particle size effect or the differences
in the textural porosity (see Tables 1 and 2) since the presence
of channels can enhance the activity compared with the lathlike
morphology in MSU-␥ catalyst. The high turnover frequency value
obtained for the largest pore 1% Pd/SBA-15 catalyst at S:Pd = 11,000
may be attributed to several factors such as (i) facile solvation of the
reactant in the mesostructured pores, which allows for an increase
in active site accessibility in comparison with the other studied sup-
ports and (ii) differences in the particle size or dispersion values
of the catalysts. Several authors [47–52] have reported differences
in the specific activity as the particle size decreases or increases,
whereas others found the TOF to increase upon increasing metal
dispersion [53]. For the hydrogenation of dienes, the specific activ- Fig. 7. Hydrogen uptake curves for the hydrogenation of 3-hexyne on Pd-supported
catalysts. T = 298 K, PH2 = 40 psig, S/Pd = 11,000.
ity is constant up to certain dispersion value between 20% and
35% depending on the support but then decreases with increasing
palladium dispersion [54,55]. This apparent disagreement may be of 3-hexyne have been reported [57]. Mesostructured solids, such
understood by considering that these reactions have been carried as MSU-␥, SBA-15 and MCM-48, display a new advantage for the
out under different experimental conditions. hydrogenation of 3-hexyne as supports of palladium catalysts. As
The synthetic potential and utility of alkynes in fine chemistry is shown in Fig. 7 the hydrogen uptake obtained for the hydrogenation
attributed to the possibility to form new carbon–carbon by alkyla- of 3-hexyne over the Pd-supported catalysts corresponds to a typi-
tion. Also, the selective hydrogenation of the triple bonds opens cal zero-order reaction, which is characteristic of alkene and alkyne
up routes to alkanes by complete saturation and to alkenes by hydrogenations over supported catalysts [5–7,62]. Even though the
semi-hydrogenation [29]. Palladium is the most efficient metal trend was similar for all studied catalysts, significant differences in
with regard to both alkene formation and syn-addition to give the catalytic activity were observed, with the 1% Pd/SBA-15 cata-
the cis-alkene [1,5–7,56]. The semi-hydrogenation of 3-hexyne has lyst being the most active. The initial rate was maintained as the
been investigated on some Pd-supported catalysts [1,57–61]. For reaction progressed, indicating that the active sites were readily
Pd/GO (graphite oxide), high selectivity to the (z)-alkene is main- accessible for the 3-hexyne molecules.
tained up to conversions of 85% at S:Pd = 10,000 for the 3-hexyne A summary of the reaction rates, expressed as the zero-
hydrogenation and a linear trend for hydrogen consumption ver- order specific rate constant and turnover frequency (TOF) at
sus reaction time have been reported [59]. For alumina supported S:Pd = 11,000, is provided in Table 4 for the three studied catalysts.
[PdCl2 (NH2 (CH2 )12 CH3 )2 ] complex, a ∼99% (z)-alkene selectivity The 1% Pd/SBA-15 catalyst proved to be a particularly efficient cat-
at 303 K and alkyne/Pd molar ratio of 8400 for the hydrogenation alyst for the hydrogenation of 3-hexyne at S:Pd = 11,000, for which

Table 4
Specific rate constant (k), turnover frequency (TOF) and selectivity for the hydrogenation of 3-hexyne on supported Pd catalysts at a substrate:Pd molar ratio of 11,000.

Catalyst k (mol L−1 min−1 g−1 Cat.) TOF (s−1 )a Selectivity (%)b

cis-3-Hexene trans-3-Hexene cis- and trans-2-hexene Hexane

1% Pd/SBA-15 4.0 2538 98.3 0.6 – 1.1


1% Pd/MCM-48 1.9 730 97.1 1.6 0.3 1.0
1% Pd/MSU-␥ 1.7 111 96.2 3.1 0.2 0.5
a
Calculated from H2 chemisorption.
b
Conversion level 70%.
Author's personal copy

G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34 33

porous Pd-supported catalysts with differences in the selectivity


levels (see Table 4). The 1% Pd/MSU-␥ and 1% Pd/MCM-48 cat-
alysts show higher selectivity towards trans-3-hexene formation
at around 70–80% conversion compared with the other catalysts.
The enhanced selectivity of trans-3-hexene formation and small
amounts of 2-hexenes isomers, observed for 1% Pd/MSU-␥ and
1% Pd/MCM-48 catalysts, may be related to the decrease of the
hydrogenation rate after approximately 30 min of reaction, as dis-
played in Fig. 7. On the other hand, the selectivity to cis-3-hexene
obtained for the commercial 1% Pd/Al2 O3 catalyst was 97% with a
low proportion of trans-3-hexene, 2-hexenes isomers and hexane,
respectively. Table 4 summarises the selectivity to the cis-3-hexene
is in the range 96–98% for conversion level 70% for all Pd-supported
catalysts. Recently Mastalir et al. [59] showed for the transforma-
tion of 3-hexyne over a Pd-graphite oxide catalyst, a high selectivity
to the cis-isomer (98%) with a small amounts of trans-isomer and
alkane formed as co-products. They attributed the improved activ-
ity and selectivity to the relative small size of the reactant, which
ensures ready access to the interlamellar active sites of the catalyst.

Fig. 8. cis-3-Hexene selectivity as a function of conversion for 3-hexyne hydrogena- 4. Conclusions


tion over 1% Pd/SBA-15 catalyst at 298 K, PH2 = 40 psig and S/Pd = 11,000.
The obtained results show that, in general, the Pd supported
hydrogenation was completed in 17 min and the stereo-selectivity on mesostructured solids are effective catalysts for the stereo-
of cis-3-hexene formation was 98.3%. selective hydrogenation of 3-hexyne to cis-3-hexene in toluene.
For further comparison, the catalytic performance of a com- The most active catalyst was 1% Pd/SBA-15 in comparison to 1%
mercial supported Pd catalyst (1% Pd/Al2 O3 ) was investigated for Pd/MSU-␥, 1% Pd/MCM-48 and commercial 1% Pd/Al2 O3 catalysts.
the same reaction at T = 298 K, PH2 = 40 psig and S/Pd = 11,000. The This behaviour can be attributed to several factors such as (i) the
results are displayed in Fig. 7. For the transformation of 3-hexyne, mesoporous framework effect which mediates substrate access to
the catalytic activity of the commercial supported Pd on alumina active sites on the palladium particles and diffusion of the cis-
catalyst proved to be close in activity (k = 2.5 mol L−1 min−1 g−1 olefin away from active sites, (ii) differences in the textural porosity
Cat.) to those of the mesoporous catalyst. Nevertheless, for 1% observed for these catalysts which also mediates substrate and
Pd/SBA-15 which proved to be a more active catalyst may be related product diffusion to and from active sites and (iii) the size of
in part to the effect of the support, which ensured a more facile the metal particles or the metal particle dispersion effect on the
dispersion of this catalyst in toluene than that of 1% Pd/Al2 O3 . catalysts. The dominant species is cis-3-hexene, which is either
The semi-hydrogenation of internal alkynes over palladium cat- hydrogenated to hexane or isomerizes to trans-3-hexene at very
alysts typically results in the predominant formation of cis-isomers high conversion, the appearance of trans-3-hexene, hexane and
accompanied by the production of alkanes, trans-isomers which 2-hexenes isomers is dependent on the support used.
may be formed either as initial products or via cis–trans isomer-
ization [5]. Fig. 8 shows the evolution of the selectivity during
References
3-hexyne hydrogenation for a representative Pd-supported cata-
lyst. The main product is cis-3-hexene with a low proportion of [1] A. Mastalir, B. Rác, Z. Király, A. Molnár, J. Mol. Catal. A: Chem. 264 (2007)
trans-3-hexene and hexane, respectively. No isomerization from 170–178.
[2] P. Han, X. Wang, X. Qiu, X. Ji, L. Gao, J. Mol. Catal. A: Chem. 272 (2007) 136–141.
3-hexene to 2-hexenes isomers was observed for the palladium
[3] I. Yuranov, P. Moeckli, E. Suvorova, P. Buffat, L. Kiwi-Minsker, A. Renken, J. Mol.
catalyst supported in SBA-15, while for the others catalysts very Catal. A: Chem. 192 (2003) 239–251.
small amounts of 2-hexenes isomers from the isomerization of cis- [4] J. Panpranot, O. Tangjitwattakorn, P. Praserthdam, J.G. Goodwing Jr., Appl. Catal.
and trans-3-hexene was observed (∼0.3%). Fig. 8 shows that the A: Gen. 292 (2005) 322–327.
[5] N. Marín-Astorga, G. Pecchi, J.L.G. Fierro, P. Reyes, Catal. Lett. 91 (2003)
selectivity for the 1% Pd/SBA-15 catalyst is fairly constant up to 115–121.
approximately 97% conversion; only a minor decrease in selectiv- [6] N. Marín-Astorga, G. Pecchi, J.L.G. Fierro, P. Reyes, J. Mol. Catal. A: Chem. 231
ity could be observed as the reaction progressed. Simultaneously, (2005) 67–74.
[7] N. Marin-Astorga, G. Pecchi, T.J. Pinnavaia, G. Alvez-Manoli, P. Reyes, J. Mol.
a small increase in the hexane selectivity is observed. The predom- Catal. A: Chem. 247 (2006) 145–152.
inant formation of cis-3-hexene is attributed to how the hydrogen [8] J. Beck, J. Vartuli, W. Roth, M. Leonowicz, C. Kresge, K. Schmitt, C. Chu, D. Olson,
is added to the substrate from active sites on the surface of the E. Sheppard, S. McCullen, J. Higgins, J. Schlenker, J. Am. Chem. Soc. 114 (1992)
10834–10843.
catalyst. The predominant formation of cis-3-hexene is attributed [9] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, US Patent 5,098,684 (1992).
to the pathway for addition of hydrogen to the adsorbed substrate [10] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck, Nature 359 (1992)
at active sites on the surface of the catalyst [29,60]. As suggested 710–712.
[11] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, US Patent 5,102,643 (1992).
by Ulan et al. [60] morphological differences in palladium par-
[12] W.J. Roth, US Patent 6,096,288 (1998).
ticles affect the selectivity of alkyne hydrogenation. Maximum [13] C. Thoelen, K. Van de Walle, I.F.J. Vankelecom, P.A. Jacobs, Chem. Commun. 18
cis-hexene selectivity (∼87%) was obtained when the substrate was (1999) 1841–1842.
[14] M. Chatterjee, Y. Ikushima, F.Y. Zhao, Catal. Lett. 82 (2004) 141–144.
adsorbed on the (1 1 1) Pd single crystal plane, whereas for adsorp-
[15] M. Thommes, R. Kohm, M. Froba, Stud. Surf. Sci. Catal. 128 (2000) 259–277.
tion on the (1 1 0) surface of the metal, the selectivity was extremely [16] P.T. Tanev, T.J. Pinnavaia, Science 267 (1995) 865–867.
unselective. [17] S.-S. Kim, T.R. Pauly, T.J. Pinnavaia, Chem. Commun. (2000) 835–836.
The pronounced (cis) stereo-selectivity of 1% Pd/SBA-15 cat- [18] S.-S. Kim, T.R. Pauly, T.J. Pinnavaia, Chem. Commun. (2000) 1661–1662.
[19] T.J. Pinnavaia, P.T. Tanev, W. Zhang, J. Wang, M. Chibwe, US Patent 6,193,943,
alyst was essentially maintained throughout the reaction (see February 27 (2001).
Fig. 8). Similar behaviour was exhibited by the other meso- [20] T.R. Pauly, T.J. Pinnavaia, Chem. Mater. 13 (2001) 987–993.
Author's personal copy

34 G. Alvez-Manoli et al. / Applied Catalysis A: General 387 (2010) 26–34

[21] B.L. Newalkar, N.V. Choudary, U.T. Turaga, R.P. Vijayalakshmi, P. Kumar, S. [41] J.M. Kim, S.K. Kim, R. Ryoo, Chem. Commun. 2 (1998) 259–260.
Komarneni, T.S.G. Bhat, Microporous Mesoporous Mater. 65 (2003) 267–276. [42] W. Zhao, Z. Hao, C. Hu, Mater. Res. Bull. 40 (2005) 1775–1780.
[22] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky, [43] S.J. Gregg, K.S.W. Sing, Adsorption, Surface Area and Porosity, Academic Press,
Science 279 (1998) 548–552. London, 1991.
[23] I. Pálinkó, Appl. Catal. A: Gen. 126 (1995) 39–49. [44] R.W. Hicks, N.B. Castagnola, Z. Zhang, T.J. Pinnavaia, C.L. Marshall, Appl. Catal.
[24] M.P.R. Spee, J. Boersma, M.D. Meijer, M.Q. Slagt, G. Van Koten, J.W. Geus, J. Org. A: Gen. 254 (2003) 311–317.
Chem. 66 (2001) 1647–1656. [45] M. Crocker, R.H.M. Herold, J.G. Buglass, P. Companje, J. Catal. 141 (1993)
[25] O. Dominguez-Quintero, S. Martinez, Y. Henriquez, L. D’Ornelas, H. Krentzien, 700–712.
J. Osuna, J. Mol. Catal. A: Chem. 197 (2003) 185–191. [46] J.A. Bennett, R.P. Fishwick, R. Spence, J. Wood, J.M. Winterbottom, S.D. Jackson,
[26] A. Corma, Chem. Rev. 97 (1997) 2373–2419. E.H. Stitt, Appl. Catal. A: Gen. 364 (2009) 57–64.
[27] R.M. Rioux, H. Song, J.D. Hoefelmeyer, P. Yang, G.A. Somorjai, J. Phys. Chem. B [47] J.P. Boitiaux, J. Cosyns, S. Vasudevan, Appl. Catal. A: Gen. 6 (1983) 41–51.
109 (2005) 2192–2202. [48] Y.A. Ryndin, L.V. Nosova, A.I. Boronin, A.I. Chuvilin, Appl. Catal. A: Gen. 42 (1988)
[28] J. Zhu, Z. Konya, V.F. Puntes, I. Kiricsi, C.X. Miao, J.W. Ager, A.P. Alivisatos, G.A. 131–141.
Somorjai, Langmuir 19 (2003) 4396–4401. [49] N. Semagina, A. Renken, L. Kiwi-Minsker, J. Phys. Chem. 111 (2007)
[29] R.L. Augustine, Heterogeneous Catalysis for the Synthetic Chemist, Marcel 13933–13937.
Dekker, Inc., New York, 1996, pp. 41–43, 387–401 (Chapters 3 and 16). [50] J.A. Anderson, J. Mellor, R.P.K. Wells, J. Catal. 261 (2009) 208–216.
[30] D. Kumar, K. Schumacher, C. du Fresne von Hohenesche, M. Grun, K.K. Unger, [51] A. Mastalir, Z. Kiraly, F. Berger, Appl. Catal. A: Gen. 269 (2004) 161–168.
Colloids Surf. A: Physicochem. Eng. Aspects 187–188 (2001) 109–116. [52] C.E. Gigola, H.R. Aduriz, P. Bodnariuk, Appl. Catal. A: Gen. 627 (1986) 133–144.
[31] A. Monier, F. Schuth, Q. Huo, D. Kumar, D. Margolese, R.S. Maxwell, G.D. Stucky, [53] A. Sarkany, A.H. Weiss, L. Guczi, J. Catal. 98 (1986) 550–553.
M. Krishnamurty, P. Petroff, A. Firouzi, M. Janicke, B.F. Chmelka, Science 261 [54] G. Deganello, D. Duca, A. Martorana, G. Fagherazzi, A. Benedetti, J. Catal. 150
(1993) 1299–1303. (1994) 127–134.
[32] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, J. Am. Chem. Soc. 120 (24) [55] L.F. Liotta, A.M. Venezia, A. Martorana, G. Deganello, J. Catal. 171 (1997)
(1998) 6024–6036. 177–183.
[33] Z. Zhang, T.J. Pinnavaia, J. Am. Chem. Soc. 124 (41) (2002) 12294–12301. [56] G. Del Angel, J.L. Benitez, J. Mol. Catal. A: Chem. 94 (1994) 409–416.
[34] Z. Zhang, R.W. Hicks, T.R. Pauly, T.J. Pinnavaia, J. Am. Chem. Soc. 124 (8) (2002) [57] D.A. Liprandi, E.A. Cagnola, M.E. Quiroga, P.C. L’Argentiere, Catal. Lett. 128
1592–1593. (2009) 423–433.
[35] S.A. Bagshaw, E. Prouzet, T.J. Pinnavaia, Science 269 (1995) 1242–1244. [58] A. Mastalir, Z. Király, M. Benkő, I. Dékány, Catal. Lett. 124 (2008) 34–38.
[36] P.T. Tanev, T.J. Pinnavaia, Science 271 (1996) 1267–1269. [59] A. Mastalir, Z. Király, A. Patzkó, I. Dékány, P. L’Argentiere, Carbon 46 (2008)
[37] S.A. Bagshaw, T.J. Pinnavaia, Angew. Chem. Int. Ed. Engl. 35 (1996) 1102–1105. 1631–1637.
[38] F.J.P. Vaudry, S. Khodabandeh, M.E. Davis, Chem. Mater. 8 (1996) 1451–1464. [60] J.G. Ulan, W.F. Maier, D.A. Smith, J. Org. Chem. 52 (1987) 3132–3142.
[39] S. Cabrera, J. El Haskouri, J. Alamo, A. Beltran, D. Beltran, S. Mendioroz, M.D. [61] S. Siegel, J.A. Hawkins, J. Org. Chem. 51 (1986) 1638–1640.
Marcos, P. Amoros, Adv. Mater. 11 (1999) 379–381. [62] G. Carturan, G. Facchin, G. Cocco, S. Enzo, G. Navazio, J. Catal. 76 (1982)
[40] F. Chen, L. Huang, Q. Li, Chem. Mater. 9 (1997) 2685–2686. 405–417.

You might also like