You are on page 1of 15

PERGAMON Carbon 38 (2000) 1241–1255

Palladium catalysts on activated carbon supports


Influence of reduction temperature, origin of the support and
pretreatments of the carbon surface
a a a, b b
M. Gurrath , T. Kuretzky , H.P. Boehm *, L.B. Okhlopkova , A.S. Lisitsyn ,
V.A. Likholobov b
a
¨ Anorganische Chemie, Universitat
Institut f ur ¨ Munchen
¨ ¨
, Butenandtstr. 5 -13 ( Haus D), D-81377 Munchen , Germany
b
Boreskov Institute of Catalysis, Novosibirsk 630090, Russia
Received 25 February 1997; accepted 21 January 2000

Abstract

A number of activated carbons produced from peat, coconut shell, and by pyrolysis of hydrocarbons have been subjected
to treatment with oxygen, chlorine, hydrogen or ammonia at elevated temperatures to get a representative series of catalyst
supports differing in porous structure and surface chemistry (characterized by nitrogen adsorption and selective titrations).
Palladium was deposited from anionic (H 2 PdCl 4 ), neutral (Pd(OAc) 2 , in acetone), and cationic ([Pd(NH 3 ) 4 ](NO 3 ) 2 )
complexes. Temperature-programmed reduction, CO chemisorption, and testing in olefin hydrogenation were used to study
the possible effects of preparation variables. The origin of the carbon support and the temperature of the catalyst reduction
with hydrogen proved to have a profound influence on the properties of the catalysts. In contrast, no unambiguous correlation
has been found between catalyst properties and the pretreatments of the carbons. It is concluded that the effect of the support
comes mainly from the differences in the porous structure of the carbons, and occurrence of steric hindrance for organic
substrates in contacting the metal particles on the microporous supports is suggested.  2000 Elsevier Science Ltd. All
rights reserved.

Keywords: A. Activated carbon, Catalyst support; D. Catalytic properties, Microporosity, Surface properties

1. Introduction preparation are concerned. Even in the case of oxide-


supported systems there is still not much scientific litera-
Activated carbons have some advantages as catalyst ture on the chemical processes taking place during catalyst
supports. They are relatively inexpensive, possess a high generation and on the factors governing the course of the
surface area, allow easy recovery of supported metal by processes. However, catalysts on activated carbons are
simple combustion of the support, show chemical inertness more difficult to investigate, firstly, because of a more
both in acidic and basic media, and at the same time do not complex porous structure, and, secondly, because of the
contain very strongly acidic centers on their surface, which possible variation in the surface chemistry of the supports.
could provoke undesirable side reactions during the cata- There is also the well-known difficulty of the application
lytic run. All that makes them attractive or even indispens- of traditional physical investigation methods, e.g. spectro-
able as supports in many cases, in particular in small-scale scopic ones, in this case due to the non-transparency of
organic syntheses catalyzed by platinum group metals carbon materials to IR, visible and UV radiation. There-
(PGM) [1]. fore, despite a great demand for new catalysts, with the
Although PGM / carbon catalysts were, in all likelihood, characteristics of the carbon support and the state of the
the first supported PGM catalysts made, as described more active phase adjusted to specific conditions of the catalyst
than a century ago, catalysts on carbon supports remained application, the preparation methods continue to be pre-
rather little studied for a long time as far as the aspects of dominantly empirical in nature. Only recently the number
of scientific publications on the subject started to show an
*Corresponding author. increasing tendency [1–4].

0008-6223 / 00 / $ – see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S0008-6223( 00 )00026-9
1242 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

The two metals which are most widely used in catalysis Somewhat surprisingly, the effect of the carbon pretreat-
by PGM / carbon are platinum and palladium. From the ment proved rather small, in particular with Pd acetate as
economics standpoint, they should be provided with the the precursor. Not all supports and catalysts were studied
highest dispersion (the fraction of atoms on the surface to the same extent but it was found that, in general, the
relative to their total number, both on the surface and in origin of the carbon used and the temperature of the
the bulk of the metal particles). Factors affecting the reductive treatment of the catalysts clearly prevailed in
obtained metal dispersion may be the pore structure of the terms of their influence on the catalyst properties.
carbon, i.e. the pore width and the surface area, and the
chemical state of the surface, e.g. its coverage with oxygen
complexes. The obtained dispersions tend to be relatively 2. Experimental
low in microporous carbons [5,6] while higher dispersions
can be obtained when mesopores are present [6]. 2.1. Reagents
The possibility of easy surface cleaning by oxidative
calcination is often perceived as an advantage of using Reagent grade chemicals were used in all cases. Nitro-
metal oxides as supports for catalytically active metals. gen (99.999%), hydrogen (99.999%), oxygen (99.995%)
However, this leaves no room for variation of the nature and carbon monoxide (99.97%) were obtained from Linde
and concentration of specific surface sites, in contrast to AG, and ammonia (99.98%) from BASF. They were used
the situation with carbons. It is assumed that functional without further purification.
groups on the support surface can act as anchoring sites for
metal complexes. The role of oxygen complexes on carbon 2.2. Carbons and their treatment
surfaces on the metal dispersion has been the subject of
´
numerous studies (see references cited by Roman-Martınez ´ The starting activated carbons were selected from
et al. [7] and Refs. [6,8–10]). Many of these aspects are commercially available samples. The main part of the
described in detail in a recent review [2]. However, most work was carried out with Anthralur STA, abbreviated as
of these studies were concerned with platinum catalysts, Ant, produced from peat by Lurgi (Frankfurt / M) and AR,
whereas fewer publications dealt with Pd / C catalysts made from coconut shells (Sutcliffe Speakman Carbon,
[8,9,11–18], and the supports used often were non-porous Ashton-in-Makerfield, UK). In some cases, Norit SGM
carbon blacks but not activated carbons. As a rule, the (Nor, from peat, by Norit, Amersfoort, The Netherlands)
previous studies concerned the effect of preparation vari- and Sibunit (Sib, mesoporous carbon, produced industrially
ables either on the state of the resulting metal or on its in Russia by hydrocarbon pyrolysis, followed by gas
catalytic properties only. In addition, relatively high tem- activation [23,24]) were used for comparison.
peratures were usually applied in the reduction of the Following grinding (if necessary), fractions,100 mm
supported phase, which is rare in real practice. were isolated by sieving. To remove ash constituents from
This paper describes a part of our study carried out with the activated carbons, especially transition metal oxides,
carbon-supported Pd and Pt catalysts (see also Refs. they were treated for 3 days with boiling concentrated
[19,20]). The focus of our attention was in the possible hydrochloric acid, then washed with hot water in a Soxhlet
influence of the chemical state of carbon surface and of the extractor for 3–4 weeks, and finally outgassed at 570 K in
palladium precursor compounds on the genesis of such vacuo (the extended washing and outgassing were neces-
catalysts, in view of the discrepancies on the subject in the sary to remove all chlorine from the activated carbons).
existing literature. To be sure that the effect of surface After such purification the ash residue did not exceed 2.6%
functional coverage, if found, is not an accidental one, for Ant, 2.1% for Nor, and 0.1–0.2% in the other cases.
several activated carbons from different sources have been Gas treatments were performed in a vertical reactor
used. They have been subjected to pretreatments in oxi- made from quartz glass. The pre-heated gas passed from
dizing or reducing surroundings, and the corresponding the bottom through a bed of the carbon (usually 6 g)
catalysts were produced with variation of the preparation resting on a porous silica frit of ca. 25-mm diameter.
conditions. Also, activated carbons are included in this To remove surface complexes, the carbons were heated
study that had undergone reaction with chlorine or am- to 1270 K in an N 2 stream for 6 h. They were exposed to
monia at elevated temperatures. It has been shown that on the ambient atmosphere only after cooling to room tem-
chlorination of carbon blacks chlorine is bound to sp 2 and perature under N 2 . Surface cleansing with saturation of
sp 3 carbon atoms [21], and nitrogen functions similar to reactive surface sites by hydrogen was achieved using an
pyrrole and pyridine are produced on the carbon surface analogous treatment under H 2 at 1170 K.
after the treatment with ammonia [22]. Particular care has Surface oxides were produced by heating the carbons
been taken to avoid side factors in the catalyst preparation, first to 650 K under N 2 ; after 15 min the gas was switched
such as the presence of adsorbed admixtures on the surface to O 2 (50 ml min 21 , the temperature in the carbon bed
of the supports or the detrimental influence of water, HCl rose to ca. 770 K) and after 30 min the oxidized carbons
and other reduction products (by means of high flow-rates were cooled under N 2 to room temperature. Using this
of the reducing gas). procedure, the burn-off could be kept below 10%. The
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1243

samples were homogenized by mechanical agitation over tration in the solutions being adjusted to get a metal
several hours, this was also done when several batches loading of 100 mmol per gram of support, in some cases
were oxidized to have a larger quantity of a sample. also 50 and 200 mmol g 21 . Samples with supported Pd
In order to fix nitrogen functions on the surface, carbons complexes were dried in an oven at |380 K, followed by
were treated at 1170 K with N 2 (1 h), then with NH 3 (4 h), final outgassing in a high vacuum at the same temperature
and cooled under N 2 . During NH 3 treatment, a substantial (420 K in some cases). A small amount of the resulting
fraction of the carbon is gasified as CH 4 , HCN and (CN) 2 sample (not more than 100 mg) was then taken for
[25] and nitrogen is bound to the surface in the form of reduction. With the exception of TPR experiments, the
various functions that can be identified using XPS reduction was performed in pure hydrogen, with a flow-
[22,26,27]. Chemisorption of chlorine was effected by rate of 20 ml min 21 and heating to the final temperature at
passing a Cl 2 stream (20 ml min 21 , 2 h) through the a heating rate of 20 K min 21 .
carbons which had been pretreated under N 2 (15 min) to
473 or 723 K. The nitrogen content of the prepared 2.5. BET surface areas and metal dispersion
samples was determined using the Kjeldahl method as
described earlier [22,25], and chlorine by fusion of the The N 2 adsorption isotherms for the determination of
carbons with Na 2 O 2 1Na 2 CO 3 in a Parr bomb, followed the BET surface areas were measured volumetrically at 77
by potentiometric titration of Cl 2 with a standard AgNO 3 K, using an apparatus built from stainless-steel tubing and
solution. Swagelok connections. Only the sample bulb was made
The gas-treatment performed with a carbon will be from glass. Pressures were measured using Baratron gages.
indicated in the code for the carbon by adding -H, -ox, -Cl Micropore volumes, Vmicro , were determined in the usual
or -am, respectively. Carbons which underwent no treat- way from Dubinin–Radushkevich (DR) plots of the N 2
ment, other than the previous HCl extraction and evacua- adsorption isotherms. The widths of the micropores, Lmicro ,
tion, are designated by the symbol -u and those heat- were estimated from the DR characteristic energy, E0 ,
treated under N 2 at 1170 K by -HT. following Stoeckli et al. [30]: Lmicro 510.8 /(E0 211.4)
The conditions of both the treatments and the storage of with E0 in kJ mol 21 and L in nm. For the surface area of
the samples (in tightly closed vessels) prevented contami- ideal, slit-shaped pores it follows from geometrical consid-
nation of the carbon surface by side-products and volatile erations that Smicro 52Vmicro /Lmicro . Mesopore volumes and
contaminants in the air. Additionally, all the carbons were surface areas were determined from the desorption branch-
outgassed at 570 K at 10 22 Pa immediately prior to their es of the isotherms with the assumption of slit-shaped
use as supports or performing titration and nitrogen pores, following the procedure outlined by Orr and Dalla
adsorption measurements. In particular, such a final treat- Valle [31].
ment can be considered adequate for getting correct Metal dispersions were measured using dynamic CO
titration data on the nature and concentration of surface chemisorption. A Pulse Microsorb 2700 instrument (Mi-
oxides (especially CO 2 and strongly retained HCl from the cromeritics) was used with hydrogen as the carrier gas
initial cleaning procedure affect the results of the acid– [12,32,33]. A freshly reduced surface was prepared by in
base titrations.) situ heating of the sample (ca. 10 mmol metal) to 323 K
under hydrogen for 1 h. These reduction conditions were
2.3. Titration measurements identical to those in the preparation of the catalysts. After
Acidic groups on the carbon surface were determined cooling to room temperature and stabilization of the heat-
from adsorption neutralization with NaHCO 3 , Na 2 CO 3 , conductivity cell, several 70-ml pulses of CO were injected
NaOH or NaOEt solutions. Three-hundred milligram sam- into the H 2 stream (20 ml min 21 ). Since CO is slowly
ples were equilibrated overnight with 25 ml of 0.05 N desorbed under flowing H 2 , consecutive pulses were
NaHCO 3 , Na 2 CO 3 or NaOH solutions; then 10 ml 0.05 N injected immediately after the curve had returned to the
HCl were added to 10 ml aliquots and back-titrated with base line (0.5–1 min). It was ascertained that the loss of
0.05 N NaOH after boiling off the CO 2 . This has been adsorbed CO was negligible under these conditions (the
described in detail elsewhere [28,29]. Adsorption from 0.1 results with a few test samples agreed well with those
N sodium ethoxide in ethanol was determined analogously. obtained under static adsorption measurements).
Basic surface sites were determined similarly by adsorp-
tion from 0.05 N HCl solution. 2.6. Temperature-programmed reduction

2.4. Catalyst preparation The carbon samples loaded with the metal salts (40–50
mg) were outgassed at 383 K for ca. 16 h and placed into a
PdCl 2 (in 1 M HCl), Pd acetate (in acetone) and tubular flow reactor equipped with temperature-pro-
[Pd(NH 3 ) 4 ](NO 3 ) 2 (aqueous solution) have been used as grammed heating. The samples remained in a stream of the
metal precursors (all Pd compounds from Degussa). Typi- reducing gas (H 2 / N 2 1:19 v / v) overnight to stabilize the
cally, they were introduced into the pores of the supports base line of a heat conductivity cell before the heating up
by incipient-wetness impregnation, with the Pd concen- to 600 K was turned on, typically with a heating rate of 10
1244 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

K min 21 and a controlled gas flow of 20 cm 3 min 21 . lytic runs were carried out upon adding a new portion of
Volatile reduction products were frozen out at 77 K in a cyclohexene into the reaction mixture after completion of
trap between the reactor and the catharometer. the previous run.

2.7. Catalytic activity


3. Results
Cyclohexene hydrogenation was performed at 303 K
and 1 bar of total pressure. A static reactor approximately 3.1. Selective titrations and nitrogen adsorption
50 cm 3 in volume was used, and intensive stirring of the
reaction mixture was provided. Cyclohexene was prepared Data on structural characteristics of the carbons after
by dehydration of cyclohexanol (Fluka) with hot concen- different treatments and on the concentration of acidic and
trated phosphoric acid and distilled twice before use. basic sites on their surface are presented in Table 1. As
Special experiments were carried out varying the catalyst shown earlier [29], relatively strongly acidic groups such
amount added (2–30 mg), the stirring rate, or using a as carboxylic acids are neutralized by NaHCO 3 , whilst
shaker-type reactor. The results showed an absence of carboxyl groups condensed to lactone or lactol functions
strong poisons in the reaction mixture and a negligible require Na 2 CO 3 for neutralization. NaOH neutralizes, in
influence of external diffusion limitations under the stan- addition, weakly acidic groups such as phenolic OH
dard conditions applied. To minimize internal diffusion groups. Sodium ethoxide, NaOEt, reacts also with reactive
limitations, all catalysts were thoroughly ground in a carbonyl groups, leading to salts of hemiacetals.
mortar to an average particle size of approximately 5 mm In full agreement with what might be expected from
prior to the catalytic runs. previous work in the field [29], treatment with oxygen
The ground catalyst was placed into the reactor and gave supports with a significantly larger number of acidic
outgassed for 5 min. The reactor was filled with hydrogen, groups, especially those capable of being titrated by the
and after 5 min methanol was added (8 ml) and the stirring weak bases (carbonates). A simultaneous decrease in
turned on. After another 5 min the reaction was started by consumption of HCl (see Table 1) is also in agreement
injecting 0.4 ml of cyclohexene. Charging the reactor with earlier observations [29]. It is very probable that some
under such conditions ensured good reproducibility of the of these new groups were created at the expense of former
catalytic measurements. The consumption of H 2 was basic sites. As expected, the reductive treatment in nitro-
followed gas-volumetrically. In some cases, several cata- gen, hydrogen, or ammonia led to the opposite trends, that

Table 1
Characteristics of the carbon supports after various treatments
DR
Carbon SBET S micro V DR
micro Lmicro Smeso Vmeso Titration using
(m 2 g 21 ) (m 2 g 21 ) (cm 3 g 21 ) (nm) (m 2 g 21 ) (cm 3 g 21 )
NaOEt NaOH Na 2 CO 3 NaHCO 3 HCl
(meq g 21 ) (meq g 21 ) (meq g 21 ) (meq g 21 ) (meq g 21 )
Ant-u 630 640 0.246 0.77 28 0.083 890 355 155 80 400
Ant-ox 650 600 0.253 0.84 28 0.078 1430 940 560 200 300
Ant-HT 660 585 0.248 0.85 – – 360 100 65 7 450
Ant-H 710 680 0.266 0.78 31 0.072 – – – – –
Ant-am a 1000 730 0.366 1.00 – – 340 160 20 0 740
Ant-Cl b 440 410 0.170 0.86 24 0.059 880 335 65 75 –
AR-u 1310 1010 0.490 0.97 10 0.051 677 315 108 8 260
AR-ox 1390 900 0.550 1.23 16 0.070 – 400 – – 180
AR-H 1370 – – – – – 148 19 0 0 350
AR-am a 1350 850 0.520 1.24 10 0.043 – 103 – – 530
AR-Cl b 1210 920 0.490 1.06 – – – 302 – – –
Nor-u 1250 |750 0.450 |1.2 40 0.130 360 135 75 35 450
Nor-ox 1220 |720 0.430 |1.2 40 0.110 860 390 225 50 390
Nor-HT 1180 |700 0.420 |1.2 40 0.120 380 100 20 4 490
Sib-u 410 – 0.170 c 4.9 d .300 0.700 – 40 10 0 350
1.28 and 0.84 mmol N g 21 in Ant-am and AR-am, respectively.
a

All values based on carbon content to take account of mass increase by chlorination; 3.18 and 1.52 mmol Cl g 21 in Ant-Cl and AR-Cl,
b

respectively.
c
Formal value obtained by application of the DR equation, without accounting for the large surface area of the mesopores.
d
Formal calculation using the Stoeckli equation applicable at 0.4,L ,2.0.
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1245

Table 2
Changes in BET surface area and micropore characteristics of Anthralur after loading with palladium complexes or treatment with the
solvents (all samples were outgassed at 570 K prior to the nitrogen adsorption measurements)
Treatment SBET Smicro Vmicro Lmicro
(m 2 g 21 ) (m 2 g 21 ) (cm 3 g 21 ) (nm)
No treatment 618 613 0.237 (0.241)c 0.77
H2O 609 600 0.235 0.77
Acetone 566 540 0.215 (0.232)c 0.80
Pd(OAc) 2 a / acetone 578 559 0.220 0.79
Pd(OAc) 2 b / acetone 574 561 0.219 (0.239)c 0.78
Aqueous HCl 562 489 0.219 (0.232)c 0.90
H 2 PdCl b4 / H 2 O 557 474 0.217 (0.235)c 0.91
Pd loading 100 mmol g 21 .
a

Pd loading 200 mmol g 21 .


b

c
Data in parentheses for the same samples treated in hydrogen at 673 K for 4 h.

is the number of basic sites increased but the concentration 670 K, after which the micropores become accessible again
of acidic ones decreased. In the case of ammonia, since the [34,35].
reagent used for the treatment is itself a base, the increase The changes in the porosity after deposition of the metal
in surface concentration of the basic groups appeared precursors was measured with the Anthralur carbon.
especially pronounced. As has been shown earlier, Samples with supported metal complexes showed a some-
pyridine- and pyrrole-like functions are created at the what lower surface area and a decreased pore volume, as
edges of the graphene layers by such treatment [22,26]. compared to the original Anthralur (Table 2). The same
Treatment with chlorine resulted in a small decrease of the was observed, however, for the carbon treated with HCl
oxygen-containing acidic sites as well. The relatively high only, or with acetone. Acetone may undergo condensation
consumption of basic reagents in this case can be assigned, reactions under the influence of basic or acidic surface
in part, to a hydrolysis of chemisorbed chlorine, since sites producing oligomeric or polymeric products, analo-
substantial quantities of chloride ions were found in the gously to the reactions known to occur on alumina [37,38]
solutions [34,35]. Some chloride is already found in or titania surfaces [39]. As for H 2 PdCl 4 and / or HCl,
solution after washing with water. obviously, they are adsorbed predominantly in micropores
In contrast to the remarkable changes in the nature and or at the entrances to such pores, thus preventing nitrogen
concentration of surface functionalities, the conditions adsorption in part of the pores. The data in the last column
used for the treatments allowed for relatively small of Table 2 show an appreciable increase in the apparent
changes in the porous structure of the carbons. An average size of the micropores for the samples treated with
appreciable increase of apparent surface area and pore HCl or HCl1H 2 PdCl 4 , which is very likely caused by
volume can be noticed only for Anthralur treated with blocking of the smallest pores by the acidic species. One
ammonia. Substantial quantities of carbon were gasified in can see in Table 2 that the micropore volumes were largely
this case and additional micropores were created [25,36]. restored after subsequent heating to 673 K in hydrogen.
A considerable decrease of the micropore volume of Under these conditions, hydrogen chloride introduced into
Anthralur after chlorination is caused by blocking of the the sample in the treatments with H 2 PdCl 4 and / or HCl
entrances to the micropores by chemisorbed chlorine was eliminated and the chlorine content became the same
atoms. They can be removed by treatment in hydrogen at as that in the initial carbon (Table 3).

Table 3
Changes of chlorine content of Anthralur-based samples after treatment in hydrogen at increasing temperature
Sample Sample Cl content (mmol g 21 )
no.a code a
Without Reduction Reduction
reduction at 523 K (2 h) at 673 K (4 h)
1 Ant 190 155 150
2 PdCl x /Ant 710 305 –
3 HCl /Ant 365 285 165
a
1, Initial carbon; 2, sample with supported palladium chloride (incipient wetness impregnation with solution in 1 N HCl, Pd loading 100
mmol g 21 ), evacuated at 423 K; 3, Anthralur treated with 1 N HCl under the same conditions as for sample 2.
1246 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

3.2. Temperature-programmed reduction samples supported on oxidized carbons than with the
samples on a reduced carbon surface. Close inspection
TPR studies have been carried out with palladium shows, however, that also in the latter case there is a slow
complexes deposited on Ant and AR carbons. All samples hydrogen consumption already at the temperature of the
showed intense TPR signals at higher temperatures, with peaks observed with the oxygen-containing samples.
the maxima being typically in the 400–500 K range, whilst Clearly, both the reduction of the palladium species and
barely any hydrogen was consumed by the parent carbons. the reaction of hydrogen with surface functional groups are
The position, width and intensity of the peaks depend responsible for the hydrogen consumption. As soon as
significantly on the pretreatment of the supports. This is some metallic palladium is formed, dissociatively adsorbed
exemplified in Fig. 1 for samples loaded with H 2 PdCl 4 . hydrogen can move by a spillover mechanism [40] over
They showed the sharpest peaks, whilst palladium acetate the support surface to sites of chemisorbed oxygen or
gave broader signals in most cases (not shown), and chlorine and reduce them in competition with the reduction
several maxima appeared with the tetrammine complex. of the remaining Pd n 1 species [10]. A decrease of acidic
There was a much larger peak size when the carbon surface groups on treatment with hydrogen of Pd- or
surface carried chemisorbed oxygen or chlorine (-u, -ox, Pt-loaded oxidized activated carbon is confirmed by the
-Cl samples) compared to the reduced surface (heat-treated selective titration and has been observed already at 420 K
in hydrogen, inert gas, or ammonia; -H, -HT, -am). The [41].
TPR peaks occurred at a lower temperature with the Sometimes, a negative peak near 350 K was also seen. It
obviously results from decomposition of the interstitial
palladium hydride phase (b-phase) [42–44] formed from
metallic palladium already generated in the preparation of
the catalysts or during the initial flushing with hydrogen at
room temperature. The negative TPR signals were much
smaller or not discernible when oxidized or chlorinated
carbons were used as the support (-u, -ox, -Cl) or when the
tetrammine complex or palladium acetate were applied,
independently of the pretreatment of the carbons.

3.3. CO chemisorption

In the present work, a chemisorption stoichiometry of


one CO molecule per surface Pd atom is assumed. It is
well known [45] that CO is adsorbed in a bridging mode
on well-defined crystal faces of large Pd crystals with a
CO / Pd s ratio of 0.5. The stoichiometry changes, however,
with decreasing particle size, because an increasing frac-
tion of the CO is linearly adsorbed, corresponding to a
stoichiometric ratio of 1 [33,46–48]. Therefore, it is
justified to assume a ratio of 1 for small particles (,5 nm),
although the calculated dispersions tend to be somewhat
underestimated for catalysts with larger Pd particles.
Krishnakutty and Vannice found that dispersion values for
Pd on carbon supports tend to be a little higher when
adsorption of CO is used rather than that of H 2 [16]. This
is explained by a contamination of the metal surface with
carbon atoms originating from the carbon support. Appar-
ently, CO adsorption is less inhibited than that of hydro-
gen, and the authors concluded that CO chemisorption is
preferable for the determination of exposed metal atoms in
the case of Pd / carbon catalysts [17]. On ZrO 2 -supported
palladium, CO reacts above |373 K to a solid solution of
C in Pd and CO 2 [49]. Continuous disproportionation
(Boudouard reaction) with deposition of carbon on the
Fig. 1. TPR curves for differently pretreated Anthralur and AR metal surface occurs at higher temperatures, i.e. .550 K.
carbons loaded with H 2 PdCl 4 (100 mmol g 21 ). *, Sample loaded The interstitial carbon was found to be very reactive,
with 100 mmol g 21 [Pd(NH 3 ) 4 ](NO 3 ) 2 . formation of CH 4 with H 2 begins at |400 K. Thus, the
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1247

Boudouard disproportionation should not interfere with the


dispersion measurements at room temperature, and the Pd
surface should be free of carbon dissolved in interstitial
sites after exposure to H 2 at temperatures .400 K.
However, in view of the uncertainties in the reaction of
CO with Pd surfaces, the measured dispersion values
should be considered as apparent dispersions. Even if the
absolute values are not entirely correct, they reflect the
correct relative order of the dispersions.
The chemisorption properties proved to be strongly
dependent on the reduction temperature. In the majority of
cases, more or less sharp maxima in the CO uptakes as a
function of the reduction temperature were seen in the
range 370–420 K, as exemplified for one of the samples in
Fig. 2(a). On the one hand, only small CO consumption
was observed after pretreatment of the samples in hydro-
gen at room temperature, and even a prolonged exposure to
hydrogen at this temperature had little effect. As Table 4
shows, an increase in the time of pretreatment up to 3 days
led to some increase in the CO uptake, but it was still
several times smaller than the value obtained for the same
sample after 1 h reduction at 423 K. Only in some cases, a
relatively high CO uptake after H 2 treatment at room
temperature has been observed. This occurred mainly with
catalysts derived from the tetrammine complex and is
illustrated in Fig. 2(b) for Norit-supported samples, for
which such a distinction was most pronounced. A plausible
reason might be that metallic palladium was already
formed to some extent by an internal redox reaction of the
tetrammine salt during drying after the impregnation step
(see Section 2.4). Reagan et al. have reported that
[Pd(NH 3 ) 4 ] 21 ions in Y zeolites are decomposed at |520
K in an inert atmosphere [50]. It has been indicated that
the reaction may take place at lower temperatures (|430
K) on carbon surfaces [51].
On the other hand, whatever the sample, there was
always a strong decrease in the apparent palladium disper-
sion upon reduction above 420 K, obviously due to
sintering of the metal. This is clearly seen in Fig. 2, and
also in Table 5. A significant decrease of the apparent Fig. 2. Apparent dispersion values (CO / Pd) as a function of the
dispersions was observed at 573 K and higher temperatures reduction temperature (treatment with H 2 for 1 h): (a) for H 2 PdCl 4
when the duration of the reduction treatment was extended on Ant-HT; (b) [Pd(NH 3 ) 4 ](NO 3 ) 2 on Nor-ox. The metal loadings
much beyond 1 h; this is in agreement with literature were 50 mmol g 21 (open symbols) and 200 mmol g 21 (filled
reports [9,12]. symbols).
In comparison to the reduction temperature, the nature
of the palladium precursor compound and the pretreatment
of carbon support had a smaller effect on the metal Table 4
dispersion, with their influence being comparable to that of CO chemisorption on a sample of palladium acetate supported on
other variables upon catalyst preparation. Table 6 lists the AR-u carbon (100 mmol g 21 ) and treated in a H 2 stream at
dispersion values for palladium on differently modified ambient temperature for different times
Ant- and AR-carbons. These catalysts were reduced at 420
Reduction time CO / Pd a
K, which seems optimal for ensuring complete reduction
but preventing significant sintering. Compared to analo- 2h 0.100
gous Ant-based samples in Table 5, the metal loading was 1 day 0.127
3 days 0.198
halved and the conditions of drying were milder to get a
a
higher dispersion of the palladium. Table 7 presents Equals 0.56 after reduction at 423 K for 1 h.
1248 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

Table 5
Influence of reduction temperature on apparent dispersion of palladium (%) in Ant-based samples (Pd 200 mmol g 21 , final drying in vacuo
at 423 K prior to reduction for 1 h)
Support (CO / Pd)3100
T Red 5423 K T Red 5573 K
Palladium chloride as metal precursor:
Ant-u 58 19
Ant-ox 73 20
Ant-HT 64 20
Ant-H 34 14
Ant-am 35 20
Palladium acetate as metal precursor:
Ant-u 30 20
Ant-ox 27 19
Ant-HT 38 15
Ant-H 28 23
Ant-am 32 27
Ammine complex as metal precursor:
Ant-u 30 17
Ant-ox 27 18
Ant-HT 38 16
Ant-H 47 18
Ant-am 52 28

Table 6 the smallest dispersions (Tables 5 and 6). Relatively low


Apparent dispersion of palladium (%) in AR- and Ant-based dispersions of platinum have been observed earlier when
catalysts reduced under mild conditions (metal loading 100 the adsorbed [Pt(NH 3 ) 4 ] 21 complex was directly reduced
mmol g 21 , drying at 383 K, reduction at 423 K for 1 h) with hydrogen [5,7]. This is assumed to be due to the
Support (CO / Pd)3100 intermediate formation of a mobile species, probably
Precursor: PdCl 2 Pd(OAc) 2 [Pd(NH 3 ) 4 ](NO 3 ) 2 [Pt(NH 3 ) 2 H 2 ], that leads to easy agglomeration. The
authors obtained higher dispersions when the Pt tetra-
Ant-u 71 61 27 mmine complex was first thermally decomposed under
Ant-ox 70 54 15 helium prior to complete reduction with hydrogen. The
Ant-H 28 45 19
palladium tetrammine complex shows an analogous be-
Ant-am 66 37 19
Ant-Cl 68 64 5 havior [52]. Such a positive effect of the preceding
AR-u 68 56 2 decomposition of the tetrammine complex is visualized in
AR-ox 55 23 16 our case as well. One can compare the data presented in
AR-H 46 15 18 Tables 5 and 6 for Ant-supported catalysts reduced at 423
AR-am 58 47 31 K; if palladium acetate serves as the metal precursor, the
AR-Cl 21 25 6 dispersions of the samples in Table 5 appear appreciably
smaller than those in Table 6, but quite the opposite is seen
for samples prepared from the tetrammine complex. As the
catalysts in Table 5 have a higher Pd loading and were
available data for Nor-supported samples differing in the subjected to outgassing at a higher temperature before
pretreatment of the support. reduction, one should expect in this case a more pro-
The use of the tetrammine complex tended to result in nounced decomposition of the supported tetrammine com-
Table 7 plex in the drying stage. A higher extent of the decomposi-
Apparent dispersion of palladium (%) on oxidized and heat- tion of the complex at a higher loading can also explain the
treated Norit (metal loading 200 mmol g 21 , final drying at 423 K, higher apparent Pd dispersion in samples with Nor-ox as
reduction at 523 K for 1 h)
the support [Fig. 2(b)] upon changing the Pd loading from
Support (CO / Pd)3100 50 to 200 mmol g 21 , irrespective of the reduction tempera-
Precursor: PdCl 2 Pd(OAc) 2 [Pd(NH 3 ) 4 ](NO 3 ) 2 ture.
The only certain effect of carbon surface treatments,
Nor-HT 22 30 33 well-reproducible upon repeated preparations, was a small-
Nor-ox 35 37 25
er dispersion of palladium on the carbons modified in
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1249

hydrogen (Ant-H and AR-H), provided that the dispersion


on the initial carbon was high. To some extent this was
found also after treatment with chlorine. In other cases, the
results proved inconclusive, being dependent on the parent
carbon and variables in the catalyst preparation. For
instance, both PdCl 2 - and, to a lower extent, Pd(OAc) 2 -
derived samples showed a higher dispersion on oxidized
Norit (Table 7). The same was noticed for Ant-based
samples in Table 5 (first column), provided that they were
prepared from PdCl 2 but not Pd(OAc) 2 . At the same time,
according to the data in Table 6, oxidation of the original
Ant and AR carbons gave rise to a negative rather than a
positive effect on the palladium dispersion.

3.4. Cyclohexene hydrogenation

Catalytic tests have been made for samples which were


reduced under mild conditions and are characterized in
Table 6. The activity data are presented in Fig. 3 as a
function of the apparent palladium dispersion. (If not
stated otherwise, we refer here and later to the catalytic
activity with respect to the mass of catalyst, in contrast to
specific or intrinsic activity which are related to surface
metal atoms only.) Also tested were catalysts supported on
Sibunit carbon which were prepared under the same
conditions as the samples on Ant and AR carbons from
Table 6. The catalytic tests for the 1%Pd / Sib-u catalysts
prepared from palladium chloride, acetate, and the tetra-
mmine complex exhibited activity values of 2300, 2300
and 1100 cm 3 H 2 per minute and gram catalyst, respective-
ly.
The catalytic activity per surface Pd atom proved to be
dependent mainly on the nature of the metal precursor and
the starting activated carbon, but not on the pretreatments
of the carbons. For catalysts with similar Pd dispersion, it
increased in the order PdCl 2 ,Pd(OAc) 2 |
[Pd(NH 3 ) 4 ](NO 3 ) 2 and Ant,AR,Sib. As for the carbon
surface treatments, our results suggest only tentatively a Fig. 3. Catalytic activity in cyclohexene hydrogenation as a
function of apparent dispersion (CO / Pd): (a) for Pd /Ant; and (b)
lower intrinsic activity of Pd(OAc) 2 -derived catalysts on
Pd /AR catalysts from Table 6. Open circles, Pd tetrammine
the untreated and Cl 2 -treated supports and an increased complex as the metal precursor; solid circles, PdCl 2 ; stars,
activity of catalysts prepared on oxidized carbons. Pd(OAc) 2 .
A more detailed analysis of the data met with complica-
tions, as the course of the kinetic curves (rate vs. time)
appeared not to be identical for all the samples. Usually, catalysts remained essentially the same, however, if ac-
the catalysts showed some deactivation with time-on- tivities displayed in the second or third runs were com-
stream (as with the Pd /AR-H catalyst in Fig. 4, data pared, i.e. at longer residence times of the catalysts in the
marked by open symbols). The decrease in rate during the reaction mixture.
first run varied from sample to sample and, in some cases, The catalytic performance was virtually the same when
even an increase of the rate was seen (as with the second reagents (cyclohexene and hydrogen) and solvent from
catalyst, Pd /AR-Cl, in Fig. 4, open symbols). For evalua- different sources or with additional purification (repeated
tion, the activity shown in the first run at 30% conversion distillations, passing through a column with activated
for cyclohexene was taken. The initial activity in this run alumina or Pd / Pt catalyst) were used. The non-identical
could differ from the presented values by up to620%, and changes in catalytic activity with time should thus be
the activities in subsequent runs (with addition of a new attributed to chemical effects (restructuring of the active
portion of cyclohexene) comprised 80–100% of the values centers) rather than to the presence of poisoning con-
determined in the first run. The order in activity of the taminants in the reaction mixture.
1250 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

Fig. 4. Examples of kinetic curves for catalysts in cyclohexene


hydrogenation. Open symbols represent first runs and filled
symbols repeated runs with the same catalysts (with addition of a
new portion of cyclohexene 5 min after completion of prior run).
Both catalysts from [Pd(NH 3 ) 4 ](NO 3 ) 2 .

Analysis of the presented data and results of special


experiments prove also that macrokinetics (diffusion limi-
tations) play a minor role, if any, in the observed differ-
ences of the samples in catalytic activity. In Fig. 5(a) one
can see data for two catalysts, Pd /Ant-ox and Pd /AR-u,
which differed in activity per gram approximately in a 1:2
ratio. In this case, an increased quantity of the Pd /Ant-ox
catalyst was loaded into the reactor to produce a reaction
rate similar to that observed with the Pd /AR-u catalyst,
and the catalytic performance in the second runs is shown
in Fig. 5(a), i.e. for catalysts which have already ap-
proached the steady state in the reaction mixture, as the Fig. 5. (a) Reaction rate in second runs as a function of cyclohex-
true dependence of the rate on conversion may be masked ene conversion; Pd /AR-u and Pd /Ant-ox as catalysts (both from
in the first runs by catalyst activation / deactivation. One PdCl 2 ). (b) Subsequent runs with the same catalysts. At the point
can see that the decline in the reaction rate with increasing of |35% conversion (indicated by arrows) an additional quantity
conversion (i.e. with decreasing concentration of the of cyclohexene (0.5 cm 3 ) was added during these runs.
organic substrate) is approximately the same for both
catalysts. The reaction rate was fairly constant at the
beginning and decreased significantly only when the second reagent (hydrogen) would be more likely, due to its
conversion reached 70–80%. Fig. 5(b) shows subsequent low solubility in common organic solvents and a positive
(third) runs with the same catalysts in which the ex- order of the reaction rate with respect to the concentration /
perimental conditions were extended. These runs were pressure of H 2 (the order is equal to 0.5 for cyclohexene
begun as usual, i.e. 0.4 ml of cyclohexene were added to hydrogenation on Pd catalysts [53]). To check the absence
the reaction mixture after consumption of all the cyclohex- of such limitations in our case, fresh PdCl 2 -derived
ene in the preceding run. However, an additional quantity samples on Anthralur and AR supports with similar metal
of cyclohexene was added in these third runs at the time dispersion have been prepared and tested in the hydro-
indicated by arrows, corresponding to 35% conversion of genation of cyclooctene (using the same technique as in
the original cyclohexene loading. It is clearly seen that cyclohexene hydrogenation but with the use of ethanol as
such an increase in the cyclohexene concentration had no the solvent [20]). The reactivity of cyclooctene is more
significant effect on the catalytic activity. than an order of magnitude lower than that of cyclohexene,
Diffusional limitations with respect to the supply of the and thus the rate of hydrogen consumption per gram of
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1251

catalyst is much smaller. If strong diffusion limitations Pt II ) or the metallic state, as has been repeatedly described
with regard to hydrogen were responsible for the observed [8,9,51,54–58]. The reductive power of the carbon de-
difference in catalytic activity in the hydrogenation of creases with increasing surface oxidation, and the extent of
cyclohexene for the catalysts on different supports, one surface oxidation depends also on the nature of the
should expect disappearance or, at least, a strong decrease oxidant, its concentration, pH and temperature; e.g.
of such differences upon hydrogenating cyclooctene. The [Pt(NH 3 ) 4 ] 21 at pH 8.5 will not be reduced by carbon
difference remained at the same level, however. The [51]. (2) Alternatively, there can be a partial hydrolysis of
activities shown by the freshly prepared Pd /Ant (67% the precursor metal complex during catalyst preparation
dispersion) and Pd /AR (|60% dispersion) catalysts were and this will result in the deposition of easily reducible
equal to 450 and 1300 cm 3 min 21 g 21 , respectively, in the oxide or hydroxide species on the support surface. Some
hydrogenation of cyclohexene, and to 12 and 38 cm 3 oxide species of palladium and platinum are known to
H 2 min 21 g 21 , respectively, in the hydrogenation of undergo reduction with hydrogen already below room
cyclooctene on the same catalysts. It should be also noted temperature [43,59–61]. Such a reaction has been reported
that the volume of mesopores, which are mainly involved even at |173 K [62]. This hydrolysis will also be
in the reagent transport to the active phase, is larger for facilitated on a reductively treated surface. Removal of
Ant (Table 1), than for AR, but only the Pd /Ant catalysts chemisorbed oxygen will raise the concentration of basic
exhibited the lowest activity, see Fig. 3(a) and (b). surface sites which favors the hydrolytic reaction as the
sites can bind the acid (HCl or HOAc) formed in the
hydrolysis and, thus, participate in the equilibrium. Forma-
4. Discussion tion of relatively large metal or metal oxide particles prior
to the reduction in hydrogen might cause a poor dispersion
4.1. Influence of carbon pretreatments of the catalysts. However, Fig. 1 shows clearly that the
intensity of the negative TPR peaks was rather small in
In contrast with what we expected at the beginning of every case, and that, therefore, only a minor part of the
this study, the results show a rather small effect of the ratio palladium chloride was converted to the metal in the
of acidic and basic surface sites on the dispersion and preliminary stages of the TPR experiments.
catalytic activity of carbon-supported palladium, and also While it has been reported that high metal dispersions
of incorporation of nitrogen or chlorine heteroatoms in the could be obtained after ion exchange with surface oxygen
carbon surface. It may only be noticed that the reductively groups [6], further studies showed disappointingly low Pt
treated carbons (-H, -am, -HT) contain some palladium in dispersions on oxidized carbon surfaces [51,57,63]. This
a metallic or easily reducible form after the deposition of was assumed to be due to the thermal decomposition of
PdCl 2 (as evidenced by a negative TPR peak in the many surface groups, namely carboxyl groups, at relatively
low-temperature region, Fig. 1), and there is a tendency for low temperatures, i.e. above 470 K. When these groups are
the palladium to show a lower dispersion on the carbons destroyed, the fixed small metal particles (or unreduced
heat-treated in hydrogen (-H in Tables 5 and 6). On the precursor) will become mobile on the surface and will
other hand, the oxidized carbons (-ox) tend to exhibit a agglomerate to larger particles [51,63].
higher intensity of the TPR peaks at elevated temperatures, Several possible reasons may be cited for the small
and to induce also a slightly higher catalytic activity. These effect of surface coverage of the carbons on the properties
trends are consistent with other data (not reported) which of the catalysts studied in the present work. First, it can
we have for Pd catalysts on other carbon supports or partly be attributed to limitations of the applied technique
prepared in somewhat different conditions. for the determination of true Pd dispersion and, secondly,
There is little doubt that the exact position and intensity to some uncertainty in the activity values, as mentioned in
of the TPR peaks in the elevated-temperature region are Section 3.4. In particular, this seems responsible for the
determined in a significant way by the reaction of hydro- scattering of experimental points seen in Fig. 3 for
gen generated by spillover with surface oxides or chlorides catalysts prepared from the same Pd salt. Thirdly, one
and, thus, not much information with regard to the should take into account the strong influence of the
palladium can be obtained from the TPR curves above 400 reduction temperature indicated by the sharp peak of Pd
K. As for the formation of metallic palladium at low dispersion vs. temperature in Fig. 2(a). In Table 6 and Fig.
temperature, the following two mechanisms seem equally 3 catalysts are compared which have been reduced at 423
possible. (1) Solutions of salts or complex compounds of K. This temperature has been chosen as the standard one,
precious metals have oxidizing properties. In consequence, as it can be considered sufficient for the reduction of at
they will react with carbon surfaces creating surface oxides least the major part of palladium, whereas a higher
in a similar way as other oxidants, e.g. nitric acid, temperature already led to substantial decreases in both CO
hypochlorite, or peroxodisulfate ions [29]. The surface uptakes and catalytic activity. However, the optimal tem-
oxidation reaction is accompanied at the same time by a perature for achieving the highest Pd dispersion is not
reduction of the precious metals to a lower valency (e.g. necessarily the same for different catalysts.
1252 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

Besides these masking factors, it seems reasonable to all other parameters and conditions used in the catalyst
suggest the occurrence of compensation phenomena. The preparation. In particular, it may be important that all
first one refers to the surface sites capable of anchoring catalysts studied in the present investigation contained
metal complexes and stabilizing the supported metal relatively small quantities of palladium (0.5–2 wt.%) and
against sintering. Even in the original form (-u), the all supports, even those designated as ‘untreated’ (-u), had
activated carbon surfaces carry a sufficiently large number undergone some aging and some sort of thermal treatment
of such sites, as Pd dispersions as high as 60–70% could before being used as supports for the palladium salts.
be obtained with proper adjustment of the other prepara-
tion conditions. The anchoring occurs by different mecha- 4.2. Influence of the porosity of the supports
nisms, however. If some of the anchoring sites are
destroyed during a treatment, the number of the remaining In contrast to the small and variable effect of the surface
ones could still be sufficient to bind the complex, at least at treatments, the catalysts showed a significant difference in
a Pd loading of 0.5–2 wt.%, as used in the present work. intrinsic activity, depending on the activated carbon from
Alternatively, the metal precursor is bound, more or less which they were prepared, i.e. Ant, AR or Sib. AR-based
effectively, by newly created groups. For instance, catalysts are 2–3 times as active as Ant-supported ones
H 2 PdCl 4 probably interacts with olefinic p-fragments of (Fig. 3), and Sib-based catalysts are superior (|1.5 times)
the surface by exchange of Cl ligands [15], but the to those supported on AR. The same order of activity has
[PdCl 4 ] 22 anion might also be bound in acidic media by also been found for Pt catalysts on these supports [19]. In
ion-exchange on protonated oxygen-containing surface view of the small effect found of the surface functional
groups. Carboxyl groups seem most suitable to bind by coverage on the catalyst properties, the differences in
ion-exchange the positively charged [Pd(NH 3 ) 4 ] 21 com- activity for catalysts on supports of different origin should
plex, in particular at relatively high pH values. If these be ascribed to differences in their porous structure. How-
groups are destroyed by heat treatment under ammonia, the ever, the porous structure of the supports could not affect
complex may be adsorbed by ligand exchange with the significantly our experimental results by way of a diffu-
newly created N-containing groups on the modified sur- sional factor, as the results of the special experiments with
face. As a consequence, the Pd dispersion can be retained cyclooctene gave strong evidence against a predominant
at the same level. Only if there is a very strong decrease in role of diffusional limitations. We have to conclude that a
the number of active sites on the carbon surface, as upon steric but not a diffusional factor is responsible for the
heat-treatment in hydrogen which saturates radical sites observed differences in activity for the catalysts supported
and generates inert CH x fragments at the expense of on the Ant, AR and Sib carbons.
olefinic groupings and acidic centers, does a drop in Pd Apparently, a part of the smallest Pd particles was
dispersion become apparent. blocked in micropores. Small metal particles in narrow
Secondly, catalysts pass through several stages during slit-shaped pores of the activated carbons may form flat
their preparation and the surface groupings act in coopera- interfaces with opposing walls of the pore, leaving only a
tion with other factors. A positive or negative influence of part of the metal surface accessible to the relatively bulky
the carbon treatment in one stage might be compensated in organic substrate in the catalytic process. The catalytic
subsequent steps. For example, the spontaneous reduction activity proved to be the lowest for palladium on Ant,
of the metal precursor upon contacting the carbon surface which contains particularly narrow pores, but the highest
or its hydrolysis on the surface (see above) might lead to in the case of Sibunit, which is distinguished by a low
formation of massive particles of a metallic or oxidic degree of microporosity and a lack of narrow micropores
phase, respectively, i.e. to a low dispersion of the pal- (Table 1). We have thus to assume that the narrower the
ladium. However, the same processes could give rise to pores are in the support, the higher is the probability of
numerous metallic or easily reducible particles capable of inaccessibility of the metal particles.
serving as metal nuclei upon further reduction of the Such a pore effect provides an easy explanation for the
samples with hydrogen. This should be of advantage for lower intrinsic activity of PdCl 2 -derived catalysts. Evident-
obtaining a high dispersion. Even the treatment of the ly, the small and more planar molecules of H 2 PdCl 4 can
carbons in hydrogen or chlorine might give rise to a penetrate deeper and into smaller pores than the more
positive effect, as the hydrophobicity of the surface bulky precursors, [Pd(NH 3 ) 4 ](NO 3 ) 2 and trimeric
produced by such treatments facilitates removal of water [Pd(OAc) 2 ] 3 (in solvents of relatively low polarity). The
during catalyst reduction, thus diminishing its detrimental probability for the pore effect must therefore be the highest
effect [64] on metal dispersion. Such a complex interplay with palladium chloride as the metal precursor. As might
of different factors prevents a clear understanding of the be expected, the catalytic activities of the PdCl 2 -derived
effect of the carbon treatments and is probably the main catalysts differed most clearly from the other ones when
reason for the discrepancies existing in the relevant Anthralur was used as the support, which is distinguished
literature. Certainly, the role of different characteristics of by the narrowest pores [Fig. 3(a) and (b)]. At the same
the carbon support must be considered in connection with time, catalysts on mesoporous Sibunit, being prepared
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1253

from PdCl 2 and Pd(OAc) 2 and having similar dispersion, decrease appeared strong even at rather moderate tempera-
showed equal catalytic activity. tures, such as 450–500 K. Heal and Mkayula [12] also
The possibility of encapsulation of the active phase in described a drop in the apparent dispersion of Pd / C
microporous supports has long been recognized but exam- catalysts from 78% to 18% (they assumed a CO asd /
ples of experimental verification are scarce. The molecular Pd surface ratio of 0.5) when the reduction was performed at
sieving effect has been detected for special types of carbon 457 K instead of 298 K. Meanwhile, many literature
supports (carbon fibers) [65] or special methods of catalyst references for carbon-supported Pd catalysts, e.g. Refs.
preparation (such as incorporation of a metal compound in [71–74], report a rather small effect of temperature on the
the carbon precursor prior to carbonization) [66–70]. The metal particle size if the treatments are performed below
data collected in the present study show the blocking to 500–600 K. Typically, a temperature of more than 700–
occur even if common activated carbons and commonly 800 K is required for a significant enlargement of the
accepted conditions of catalyst preparation are used. particles within short times.
It is sometimes assumed that little metal is deposited in A strong decrease in CO uptakes after reductive treat-
the micropores of activated carbons because they are more ment of our Pd / C catalysts at temperatures slightly above
or less inaccessible to the typical metal complexes used as the optimal one took place independently of which acti-
precursors. However, this assumption is obviously not vated carbon was used as the support and which Pd
valid when the pores are slit-shaped and the complexes are complex was used as the metal precursor. Such a behavior
planar. Besides, there is good opportunity for the metal to is quite different from that of analogous Pt / C samples
penetrate into the micropores during thermal treatments of which showed stability in the same temperature range (at
the samples, when the complexes begin to lose their least at 420–620 K [19]). Such a difference deserves
ligands, especially in the reduction step. special study. Perhaps, the drop in apparent dispersion for
palladium at 450–500 K is caused not only by sintering of
4.3. Effect of the reduction temperature the metal but also by other reasons, e.g. deeper penetration
of palladium into the pores (so that part of the metal
TPR data published in the literature give evidence for surface becomes inaccessible not only to the organic
rather a broad range of the conditions necessary for the molecules but also to CO) or formation of surface or even
reduction to the metallic state of supported metal com- bulk carbides (due to an interaction between palladium and
pounds [61]. There are examples where the reduction of the amorphous part of the carbon support).
carbon-supported platinum metals with hydrogen is per- Whatever the reasons, such a strong effect of tempera-
formed at 570 K and higher temperatures [5,16], but other ture on the properties of Pd / C catalysts calls for extreme
authors consider treatment even at ambient temperature to caution when using these catalytic systems and choosing
be sufficient to obtain the metallic state (see Ref. [12]). methods for their investigation (e.g. chemisorption mea-
Palladium is an easily reducible metal and it has been surements in a static system with preliminary high-tem-
reported [60] that palladium oxide is reduced already at perature outgassing of the samples). This lower thermal
|293 K. However, the required temperature is dependent stability of palladium is reflected in the poorer knowledge
on several factors, in particular, on the nature of the metal on Pd / C catalysts as compared to Pt / C. It might also
complex and its dispersion state. Platinum oxide was explain why carbon-supported palladium catalysts, while
reduced at a substantially lower temperature than the demonstrating excellent properties in hydrogenation re-
chloride complex or a partially hydrolyzed chloride com- actions and other processes relating to fine organic syn-
plex [60]. Thus, the reduction at room temperature may be thesis, did not find application in high-temperature pro-
limited and the fraction of the surface metal atoms cesses. Although the route into this field is not closed for
underestimated by CO chemisorption. Indeed, the catalysts Pd / C catalysts, special carbon supports must probably be
in Fig. 2(a) are characterized by negligible CO uptakes if developed for such applications.
pretreated in H 2 at ambient temperature (see also Table 4).
The necessity of treating the samples at an elevated
temperature before CO chemisorption is consistent with 5. Conclusions
the results of the TPR study which always showed
hydrogen consumption at elevated temperatures (Fig. 1).
The somewhat lower temperature for the peak maxima in 1. Treatment of activated carbons with oxygen, hydrogen,
CO chemisorption can hardly be surprising; in this case, ammonia, chlorine or by heating them in an inert gas
the samples have been pretreated for 1 h at each tempera- provides a means of smoothly varying the surface
ture and the pretreatment was made in pure hydrogen, chemistry of carbon supports. The conditions of the
whereas the TPR technique required for a high dilution of treatment can be so adjusted that the accompanying
hydrogen with an inert gas and a temperature ramp. changes in the porosity of the supports are reduced to a
Reduction at higher temperatures provokes a decrease in minimum. The influence of the state of the carbon
the Pd dispersion, however. It attracts attention that the surface, e.g. its coverage with functional groups, on the
1254 M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255

dispersion of the active phase proved comparable, C, Guerrero-Ruiz A, Lopez-Gonzalez JD. J Catal
however, to the effect of the nature of the metal 1986;99(1):171–83.
precursor and other conditions used during catalyst [6] Prado-Burguete C, Linares-Solano A, Rodriguez-Reinoso F,
´
Salinas-Martınez de Lecea C. J Catal 1989;115(1):98–106.
preparation.
[7] ´
Roman-Martınez´ MC, Cazorla-Amoros ´ D, Linares-Solano A,
2. The capability of surface Pd species to chemisorb
´
Salinas-Martınez de Lecea C, Yamashita H, Anpo M. Carbon
carbon monoxide is a non-trivial function of the reduc- 1995;33(1):3–13.
tion temperature. Pretreatment of the supported metal [8] Suh DJ, Park T-J, Ihm S-K. Ind Eng Chem Res
precursor in hydrogen at ambient conditions does not 1992;31(8):1849–56.
ensure reduction of all palladium species and results [9] Suh DJ, Park T-J, Ihm S-K. Carbon 1993;31(3):427–35.
usually in negligible CO uptakes. Quite a slight increase [10] ´
Guerrero-Ruiz A, Badenes P, Rodrıguez-Ramos I. Appl
in the reduction temperature appears sufficient, how- Catal, A (General) 1998;173(2):313–21.
ever, to produce the highest apparent dispersions. A [11] Morikawa K, Shirasaki J, Okada M. In: Eley DD, Pines H,
further increase in the reduction temperature, even a Weisz PB, editors, Advances in catalysis and related sub-
relatively small one, has a very strong detrimental effect jects, vol. 20, New York: Academic Press, 1969, pp. 97–133.
on the accessible metal surface. Tentatively, this can be [12] Heal GR, Mkayula LL. Carbon 1988;26(6):815–23.
[13] Gurevich SV, Simonov PA, Lisitsyn AS, Likholobov VA,
ascribed to sintering of the supported metal particles but
Moroz EM, Chuvilin AL et al. React Kinet Catal Lett
also to their penetration into narrow pores and / or
1990;41(1):211–6.
coverage of the metal surface with a carbonaceous [14] Wunder RW, Cobes JW, Phillips J, Radovic LR, Peinado
contamination by interaction with amorphous parts of AJL, Carasco-Marın ´ F. Langmuir 1993;9(4):984–92.
the carbon or surface functional groups. The optimal [15] Simonov PA, Romanenko AV, Prosvirin IP, Moroz EM,
reduction temperature proved to be near 400 K with the Boronin AI, Chuvilin AL et al. Carbon 1997;35(1):73–82.
metal precursor compounds and conditions of catalyst [16] Krishnankutty N, Vannice MA. J Catal 1995;155(2):312–26.
preparation used in this investigation. [17] Krishnankutty N, Vannice MA. J Catal 1995;155(2):327–35.
3. The microporosity of traditional activated carbons [18] Krishnankutty N, Li J, Vannice MA. Appl Catal, A (General)
should be regarded as a disadvantage rather than a merit 1998;173(2):137–44.
in their use as catalyst supports. According to the [19] Okhlopkova LB, Lisitsyn AS, Likholobov VA, Gurrath M,
results obtained, quite a high dispersion of the metal Boehm HP. Appl Catal, A (General), in press.
[20] Okhlopkova LB, Lisitsyn AS, Boehm HP, Likholobov VA.
proves achievable on such supports but does not
React Kinet Catal Lett, (submitted).
guarantee a proportionally high catalytic activity. While
[21] Papirer E, Lacroix R, Donnet J-B, Nanse´ G, Fioux P. Carbon
being accessible to small molecules of carbon monox- 1995;33(1):63–72.
ide, the active sites localized in narrow pores are not [22] ¨ B, Boehm HP, Schlogl
Stohr ¨ R. Carbon 1991;29(6):707–20.
easily accessible to more bulky molecules of organic [23] Likholobov VA, Fenelonov VB, Okkel LG, Goncharova OV,
substrates, and the steric hindrance for the organic Avdeeva LB, Zaikovskii VI et al. React Kinet Catal Lett
substrates to contact the active sites means that part of 1995;54(2):381–411.
the supported metal is lost for catalysis. [24] Surovikin VF, Plaksin GV, Semikolenov VA, Likholobov VA,
Tiunova Iyu. US Patent No. 4978649, 1989.
[25] Boehm HP, Mair C, Stohr ¨ T, de Rincon ´ AR, Tereczki B.
Fuel 1984;63(8):1061–3.
Acknowledgements [26] Pels JR, Kapteijn F, Moulijn JA, Zhu Q, Thomas KM.
Carbon 1995;33(11):1641–53.
Financial support from Deutsche Forschungsgemein- [27] Matzner S, Boehm HP. Extended abstracts, 22nd Biennial
schaft and Fonds der Chemischen Industrie is gratefully Conf. on Carbon, UC San Diego, San Diego, CA, American
acknowledged. Carbon Society, 1995:600–601.
[28] Boehm HP, Diehl E. Z Elektrochem, Ber Bunsenges Phys
Chem 1962;66:642–7.
[29] Boehm HP. Carbon 1994;32(5):759–69.
References [30] Stoeckli HF, Rebstein P, Ballerini L. Carbon
1990;28(6):907–9.
[1] Auer E, Freund A, Pietsch J, Tacke T. Appl Catal, A [31] Orr C, Dalla Valle JM. In: Fine particle measurements, New
(General) 1998;173(2):259–71. York: Macmillan, 1959, pp. 271–3.
[2] Radovic LR, Rodriguez-Reinoso F. In: Thrower PA, editor, [32] Bracey JD, Burch R. J Catal 1984;86(2):384–91.
Chemistry and physics of carbon, vol. 25, New York: [33] Scholten JJF, Pijpers AP, Hustings AML. In: Bell AT,
Dekker, 1996, pp. 243–358. Carberry JJ, editors, Catalysis reviews science engineering,
[3] Radovic LR, Sudhakar C. In: Marsh H, Heintz E, Rodriguez- vol. 27, New York: Dekker, 1985, pp. 151–206.
Reinoso F, editors, Introduction to carbon technology, [34] Gurrath M, Boehm HO. Extended abstracts, 21st Biennial
Alicante, Spain: University of Alicante, 1997, pp. 103–65. Conf on Carbon, SUNY, Buffalo, NY, American Carbon
[4] Rodriguez-Reinoso F. Carbon 1998;36(3):159–75. Society, 1993:462–463.
[5] Rodriguez-Reinoso F, Rodriguez-Ramos I, Moreno-Castilla [35] Gurrath M. Untersuchungen an Palladium / Kohlenstoff-
M. Gurrath et al. / Carbon 38 (2000) 1241 – 1255 1255

¨
Katalysatoren: Einflub der Oberflachen-Modifizierung des [57] van Dam HE, van Bekkum H. J Catal 1991;131(2):335–49.
¨
Tragers und der eingesetzten Palladiumverbindungen, Dr. [58] Shaikhutdinov SK, Koshubey DL. Catal Lett
Rer. Nat. Thesis, Ludwig-Maximilians Universitat, ¨ 1994;28(2):343–50.
¨
Munchen, Germany, 1997. [59] Lieske H, Lietz G, Spindler H, Volter ¨ J. J Catal
[36] Tereczki B, Kurth R, Boehm HP. Preprints, Carbon ’80, Intl. 1983;81(1):8–16.
Carbon Conf., Baden-Baden, Germany, Deutsche Keram- [60] Pinna F, Signoretto M, Strukul G, Benedetti A, Malentacchi
ische Gesellschaft, 1980:218–222. M, Pernicone N. J Catal 1995;155(1):166–9.
[37] Lawrence BM, Hogg JW, Terhune SJ. J Chromatogr [61] Jones A, McNicol B. Temperature-programmed reduction for
1969;42(2):261–2. solid materials characterization, New York: Dekker, 1986.
[38] Seebald HJ, Schunack W. Arch Pharmazie 1972;305(6):406– [62] Attwood PA, McNicol BD, Short RT. J Catal
17. 1981;67(2):257–95.
[39] Luo SC, Falconer JL. J Catal 1999;185(2):393–407. ´
[63] Coloma F, Sepulveda-Escribano A, Fierro JLG, Rodriguez-
[40] Conner Jr. WC, Falconer JL. Chem Rev 1995;95(3):759–88. Reinoso F. Appl Catal, A (General) 1997;150(1):165–83.
[41] Kuretzky T, Boehm HP. Proc Carbon’92, Intl Carbon Conf, [64] Barbier J, Bahloul D, Marecot P. J Catal 1992;137(2):377–
Essen, Germany, Deutsche Keramische Gesellschaft, 84.
1992:260–2. [65] Jin H, Park S-E, Lee JM, Ryu SK. Extended abstracts, 21st
[42] Subramanian S. Platinum Metals Rev 1992;36(2):98–103. Biennial Conf on Carbon, SUNY, Buffalo, NY, American
[43] Chang T-C, Chen J-J, Yeh C-T. J Catal 1985;96(1):51–7. Carbon Society, 1993:561–562.
[44] Huang Y-J, Xue J, Schwarz JA. J Catal 1988;111(1):59–66. [66] Trimm DL, Cooper BJ. J Chem Soc, Chem Commun
[45] Miranda R, Wandelt K, Rieger D, Schnell RD. Surf Sci 1970(8):477–8.
1984;139(2–3):430–42. [67] Schmitt Jr. JL, Walker Jr. PL. Carbon 1971;9(6):791–6.
[46] Dorling TA, Moss RL. J Catal 1967;7(4):378–85. [68] Trimm DL, Cooper BJ. J Catal 1973;31(2):287–92.
[47] Evans J, Hayden BE, Lu G. Surf Sci 1998;360(1–3):61–73. [69] Schueller OJA, Pocard NL, Huston ME, Spontak RJ, Neenan
[48] Shen LL, Karpinski Z, Sachtler WMH. J Phys Chem TX, Callstrom MR. Chem Mater 1993;5(1):11–3.
1989;93(12):4890–4. [70] Hutton HD, Pocard NL, Alsmeyer DC, Schueller OJA,
[49] Maciejewski M, Baiker A. Pure Appl Chem Spontak RJ, Huston ME et al. Chem Mater
1995;67(11):1879–84. 1993;5(12):1727–38.
[50] Reagan WJ, Chester AW, Kerr GT. J Catal 1981;69(1):89– [71] Semikolenov VA, Lavrenko SP, Zaikovskii VI, Boronin AI.
100. React Kinet Catal Lett 1993;51(2):517–24.
[51] ´
Sepulveda-Escribano A, Coloma F, Rodriguez-Reinoso F. [72] Semikolenov VA, Lavrenko SP, Zaikovskii VI. React Kinet
Appl Catal, A (General) 1998;173(2):247–57. Catal Lett 1993;51(2):507–15.
[52] Zou WQ, Gonzalez RD. Catal Lett 1992;12(1–3):73–86. [73] Ryndin YuA, Stenin MV, Boronin AI, Bukhtyarov VI,
[53] Gonzo EE, Boudart M. J Catal 1979;52(3):462–71. Zaikovskii VI. Appl Catal 1989;54(3):277–88.
[54] Puri BR, Singh S, Mahadjan OP. Indian J Chem [74] Lamber R, Jaeger N, Schulz-Ekloff G. Surf Sci 1990;227(1–
1965;3(2):54–7. 2):15–23.
[55] Fu R, Zeng H, Lu Y. Carbon 1994;32:593.
[56] Fu R, Zeng H, Lu Y, Chan WH, Ng CF. Carbon
1995;33(5):657–61.

You might also like