You are on page 1of 8

Journal of Catalysis 298 (2013) 10–17

Contents lists available at SciVerse ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Mechanistic studies of methanol synthesis over Cu from CO/CO2/H2/H2O


mixtures: The source of C in methanol and the role of water
Y. Yang a,c, C.A. Mims b,⇑, D.H. Mei a, C.H.F. Peden a, C.T. Campbell c
a
Institute for Integrated Catalysis, Pacific Northwest National Laboratory, Richland, WA 99354, United States
b
Department of Chemical Engineering, University of Toronto, Toronto, ON, Canada M5S 3E5
c
University of Washington, Department of Chemistry, Seattle, WA 98195, United States

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

Article history: The low temperature (403–453 K) conversions of CO/hydrogen and CO2/hydrogen mixtures (6 bar total
Received 1 August 2012 pressure) to methanol over copper catalysts are both assisted by the presence of small amounts of water
Revised 18 October 2012 (mole fraction 0.04–0.5%). For CO2/hydrogen reaction mixtures, the water product from both methanol
Accepted 29 October 2012
synthesis and reverse water–gas shift serves to initiate both reactions in an autocatalytic manner. In the
Available online 12 December 2012
case of CO/D2 mixtures, very little methanol is produced until small amounts of water are added. The
effect of water on methanol production is more immediate than in CO2/D2, yet the steady-state rates
Keywords:
are similar. Tracer experiments in 13CO/12CO2/hydrogen (with or without added water) show that the
Methanol synthesis
Water–gas shift
dominant source of C in the methanol product gradually shifts from CO2 to CO as the temperature is
Copper lowered. Cu-bound formate, the major IR visible surface species under CO2/hydrogen, is not visible in
Isotope tracing CO/moist hydrogen. Though formate is visible in the tracer experiments, the symmetric stretch is absent.
Water promotion These results, in conjunction with recent DFT calculations on Cu(1 1 1), point to carboxyl as a common
Autocatalysis intermediate for both methanol synthesis and reverse water–gas shift, with formate playing a spectator
Carboxyl intermediate co-adsorbate role.
Formate Ó 2012 Elsevier Inc. All rights reserved.
Carbon source
CO

1. Introduction the carbon source in methanol synthesis still lives on. . .’’. DFT cal-
culations [31–34] suggest that the water–gas shift (WGS) mecha-
Methanol synthesis on copper-based catalysts has been a target nism on copper (1 1 1) proceeds through a carboxyl (HOCO)
of decades of research because of the significance of this reaction in intermediate rather than a formate (HCOO) species or Campbell’s
the chemical industry and methanol’s potential as a liquid energy/ proposed surface redox mechanism [35]. The mechanism of meth-
hydrogen carrier [1–7]. More active methanol synthesis catalysts, anol synthesis, however, has generally been proposed to proceed
by allowing lower temperature operation, would provide economic via CO2 conversion to a surface formate species which is then
benefits through higher conversion. The catalytic activity is deter- hydrogenated to form methanol [25], although this role of formate
mined by the mechanism of the reaction which, though heavily in methanol synthesis is also still a subject of debate. The DFT cal-
studied [8–27], is still not fully understood. A recent paper by culations in Refs. [24,25], the latest and most extensive of such,
Grabow and Mavrikakis [25] contains a thorough review of the cur- continue to support a formate pathway for methanol synthesis
rent state of our understanding and indicates that the details of the from CO2. The effects of co-adsorbates, particularly CO, were con-
mechanism are more complex than previously thought. Tracer sidered in these studies and important enhancements were pre-
experiments have shown that that CO2 is the preferred reactant dicted on surfaces crowded by CO. Despite the purported role of
over CO when H2/CO/CO2 mixtures are reacted under commercial formate and some previous evidence in TPR experiments that
conditions [28–30]. As remarked in Ref. [25] however, ‘‘since Cu methanol is made by hydrogenation of formate [21], we have re-
is also an excellent WGS (water–gas shift) catalyst, facilitating cently reported that simple surface reaction (‘‘titration’’ by H2)
the conversion of CO to CO2 and vice versa, the controversy about experiments of formate-containing adlayers (produced both by
CO2–H2 reaction and from formic acid adsorption) on copper (both
⇑ Corresponding author. Address: Department of Chemical Engineering and supported on SiO2 and unsupported) fail to produce methanol in
Applied Chemistry, University of Toronto, 200 College Street, Toronto, Ontario, hydrogen atmospheres [36]. These results rule out a simple hydro-
Canada M5S 3E5. Fax: +1 416 978 8605. genation pathway for formate at moderate surface coverages and
E-mail address: charles.mims@utoronto.ca (C.A. Mims).

0021-9517/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcat.2012.10.028
Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17 11

low temperatures, although the effects of surface crowding at determined by the standard N2O titration method [41–43]. XPS
higher coverages could possibly increase the rate of this pathway analysis of the catalyst showed no inorganic surface contaminants.
as suggested in Ref. [25]. Also of relevance for this discussion is The TKA reactor was loaded with 35 mg of this catalyst, 4 mg of
that the calculations in reference [25] also support a simultaneous which was pressed onto a tungsten mesh centered in the optical
pathway for direct conversion of CO to methanol by direct hydro- path of the transmission FTIR spectrometer. An unsupported cop-
genation via a formyl intermediate. per catalyst was prepared from ultra-pure copper powder (Alfa Ae-
Here, we provide new insights into the reaction mechanism sar spherical powder, 4 lm average diameter). Four grams of this
through transient isotope tracing to determine whether the C in catalyst were gently loaded, without pressing, into a 3 cm3 tubular
the methanol product comes predominantly from CO or CO2, and reactor made of 8 mm diameter 316 stainless steel. Overnight
through transient water additions to quantify the response time reduction in H2 at low temperatures (453 K) produced a catalyst
for water promotion to products and how it differs with CO2 versus with sufficient methanol synthesis activity for our experiments
CO-containing feed. while minimizing sintering. This reactor was not equipped with
Among possible co-adsorbate effects, water has received little FTIR windows but otherwise utilized all the gas flow switching
previous attention. Of the steady-state rate laws in Ref. [1], only and mass spectrometer (MS) capabilities of the TKA apparatus.
one has water vapor as a kinetically significant species. Earlier The N2O titration method showed 1.3 lmol Cu sites/g. Slow Cu
studies at commercial conditions have shown important effects site loss was observed over time, probably due to sintering. How-
of water and CO2 on the steady-state production of methanol from ever, over the time scales reported here, the loss was less than 20%.
CO/H2 [37,38]. In these studies, the methanol synthesis rates in
H2/CO exhibit maxima as low partial pressures of CO2 or H2O are
2.2. Kinetics measurements
included in the feed. This and similar effects have been variously
attributed to production of Cu+1 sites, management of the oxida-
Reactant gas mixtures were prepared in flow-controlled mani-
tion state of the ZnO support, and in the case of water, the
folds. In all experiments reported here, 10 sccm of total gas flow
above-mentioned production of CO2 (presumed to be preferred
was used with total pressures between 2.5 and 6.0 bar. Water
reactant) via the water–gas shift reaction. No optimum effect of
was added to flowing gas streams via a syringe pump. Deuterium
water has been noted in the rates of CO2 hydrogenation to metha-
isotopic hydrogen was used in most of the experiments since the
nol, rather an inhibition has been noted in the steady-state rate
fragmentation pattern of D4-methanol (fragments separated by
[37]. Most recent microkinetic models do not involve water or
2 m/e) makes 12C/13C isotopic tracer experiments easier to inter-
water-derived OH intermediates in the CO2 hydrogenation steps
pret. The infrared absorption (IR) spectra were obtained with a
in methanol synthesis mechanism, including the mechanism in
Bruker IFS/66S FTIR and gas analysis achieved with an Extrel
Refs. [24,25]. The role of water is confined to dissociation, re-asso-
MAX 300 mass spectrometer – useful signal/noise in the surface
ciation of OH(ads) and H(ads) and steps involving OH (ads), includ-
IR spectra required 20 s of accumulation. MS sensitivities and
ing the following WGS steps: OH(ads) + CO(ads) = HOCO(ads) and
fragmentation patterns for methanol were measured by injecting
OH(ads) + HOCO(ads) = H2O + CO2 [34]. The reverse of the latter
controlled amounts of methanol (CD3OD or CH3OH) with the
of these two steps implies H2O activation of CO2, a reaction rele-
syringe pump. The experiments described here involve simply
vant to the reverse water–gas shift (RWGS) reaction and the pres-
monitoring the transient behavior of the gas-phase products by
ent study. Our recent DFT study [26], however, indicates that water
MS, as well as the surface IR (for the supported catalyst in the
can assist in certain critical surface hydrogenation steps on
TKA apparatus) as the catalyst achieves steady state in various
Cu(1 1 1) which could have a strong enabling effect on a methanol
gas mixtures. Prior to each experiment, the catalyst was freshly
synthesis pathway. From CO2, the formation of a carboxyl (HOCO)
reduced and IR baseline spectra were obtained.
intermediate was predicted in that study to be considerably en-
For the isotope tracing experiments in 12CO2/13CO/D2 mixtures,
hanced by the presence of water, while the formate hydrogenation
the experiments began with a period under He/12CO2/D2 (1:1:6)
pathway was, in fact, predicted to be suppressed. The significant
until steady state was achieved, whereupon 13CO was substituted
mechanistic step identified in Ref. [26] involves a H2O–H(ads)
for the He. The intensities of mass fragments m/e 34–37 were used
complex which can easily hydrogenate (0.25 eV barrier) one of
to measure the content of the methanol product: odd mass frag-
the oxygen atoms in CO2 to form the carboxyl intermediate – a step
ments (35,37) arise from 13CD3OD (from 13CO) and even masses
which is predicted to be difficult (1.23 eV barrier) for adsorbed H
(34,36) from 12CD3OD (from 12CO2). A mass spectrometric artifact
alone. These recent calculations were undertaken as a result of
results in small amounts (<10%) of the odd mass m/e = 35 appear-
the experimental findings presented here which show that small
ing in the 12CO2/D2 methanol product. This was attributed to H–D
amounts of water have an important effect on the methanol syn-
exchange on the catalyst support and/or the walls of the mass
thesis reaction.
spectrometer, giving small amounts of 12CD3OH. Correction for this
exchange was made in the calculation of the methanol isotopic
2. Experimental composition.

The same catalysts and transient kinetics analysis (TKA) reactor


3. Results and discussion
system described in Refs. [39,40] were used in this study. As such,
these experimental details are only briefly summarized here.
3.1. Water effects in CO2/D2 mixtures

2.1. Catalyst preparation and characterization Fig. 1 shows initial transients in the production of methanol
(m/e fragment 31) from the freshly reduced unsupported Cu cata-
Both supported and unsupported copper catalysts were used. lyst after being exposed at 413 K to flowing 6 bar D2/CO2 = 3:1 with
The supported catalyst, 10 wt.% Cu/SiO2, was prepared by incipient and without various amounts of D2O. In the absence of any added
wetness impregnation of acid-washed Davison 645 silica using water, a lengthy induction period of approximately 1.5 h (after
aqueous solutions of Cu(NO3)2 hydrate (Aldrich, 98+%). Following switching from D2/He = 3:1) is required before the methanol level
air calcination at 653 K and reduction in pure flowing H2 at in the product achieves a steady value (curve a). The steady-state
553 K for 2 h, the catalyst contained 100 lmol Cu sites/g as methanol synthesis rate eventually reached in this instance was
12 Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17

H2 H2 + CO2 D2 D2 + CO2

Methanol (m/e 34) Intensity


c
c
Methanol (m/e=31) intensity

b
d

b
a
0 500 1000
Time / sec
0 4000 8000
Time / sec Fig. 2. Mass spectrometer intensity at m/e 34 (CD3O+) versus time during methanol
synthesis on a pre-reduced, silica-supported copper catalyst at 413 K. Reactant gas
Fig. 1. Mass spectrometer intensity at m/e 31 (CH3O+) versus time during methanol composition was initially a D2/He mixture and switched to a D2/CO2 reactant gas
synthesis over reduced metallic copper powder at 413 K, after switching from a mixture (3:1 M ratio at 6 bar in both), to which various water amounts were added:
H2/He mixture to a H2/CO2 reactant gas mixture (3:1 M ratio at 6 bar in both), with (a) no water, (b) 0.1 mol%, (c) 1 mol%.
(a) no water, and to which water was added at the following levels: (b) 0.04 mol%,
(c) 0.1 mol% and (d) 0.2 mol%.

Fig. 2 shows results obtained in the TKA with the silica-sup-


3.4 ± 1.0  10 6 lmol s 1 g 1. Based on the initial site count with ported Cu catalyst during similar experiments to those shown in
N2O titration (1.3 lmol g 1), this corresponds to a turnover fre- Fig. 1. The transient methanol synthesis rates are shown after
quency of approximately 2.6 ± 0.6  10 6 s 1. The CO2 conversion switching from D2/He to dry D2/CO2 (3:1, 6 bar) (curve a), D2/CO2
to methanol under these conditions is very low, 5  10 6. to which a 1.0  10 3 mol fraction (curve b) and D2/CO2 to which
Although this temperature is much lower than that used commer- a 1.0  10 2 mol fraction of water vapor have been added to the
cially, an Arrhenius plot of the rates at those higher temperatures feed (curve c). The addition of water to the feed (curves b and c)
continues smoothly down to this temperature [40], indicating that shortens the time required before methanol production reaches
the same mechanism and rate-controlling step(s) apply here. Re- steady state, similar to the effect seen on the unsupported catalyst;
verse water–gas shift (RWGS) is 10 times faster than methanol however, the induction periods seen on the supported copper cat-
synthesis at temperatures somewhat higher than the present study alyst are significantly shorter. The faster overall response seen on
[40]. Assuming a factor of 10 holds at 413 K, the water vapor mole the supported catalyst may arise from the hold up of water on
fraction produced by RWGS at steady state is roughly 1.5  10 4. the silica support in the form of hydroxyl groups, or by the differ-
These low values are far from RWGS equilibrium, ensuring kinetic ent distribution of site-types including adlineated support-metal
control of both processes. Interestingly water vapor production, sites, on the more highly dispersed supported catalyst. It can also
and thus the RWGS reaction, also shows an induction delay similar be seen that, similar to the unsupported catalyst, the eventual stea-
to methanol synthesis under these conditions. This induction delay dy-state rate appears to decrease as the water mol fraction in the
thus involves both methanol synthesis and RWGS. reactant mixtures increases. The production of methanol on our
When small amounts of water vapor, similar to those gener- freshly reduced Cu/SiO2 catalyst reaches a steady-state rate of
ated at steady state, are co-fed with the CO2/D2, the methanol 3.6 ± 0.4  10 4 lmol s 1 g 1 after approximately 20 min exposure
induction period is shortened considerably as shown by curves to 6 bar D2/CO2 (3:1) at 413 K. This corresponds to a turnover fre-
1b–d. These experiments, starting from freshly reduced catalysts, quency of 3.6 ± 0.8  10 6 s 1, within the error bars of that on the
show that increasing amounts of water vapor (mol fractions unsupported catalyst. Deuterium was used here to simplify the IR
4.0  10 4, 1.0  10 3, 2.0  10 3) progressively shorten the spectroscopy. A mild (1.1) positive H/D kinetic isotope effect on
induction period before steady methanol synthesis is achieved. methanol turnover frequency was previously reported for this
In these experiments where water is added to the reactants, the reaction under these conditions [40]. The similarities between
water appears in the product stream promptly, substantially rates and behaviors on the supported and unsupported catalysts
ahead of the methanol transient. The synthesis of methanol from indicate that these observations are dominated by copper metal
CO2 thus appears to require the presence of water vapor or water- chemistry. Previous studies have shown that rates on single crystal
derived intermediates on the surface. In the absence of water in copper are similar to those measured on the best Cu/ZnO catalyst
the feed, reverse water gas-shift supplies this water. Furthermore, [20] – also implying dominant metal chemistry.
since the water product from RWGS is delayed along with the The catalytic rates under these conditions are so low that con-
methanol without water in the reactant gas, it appears that the siderable time could be required before steady-state coverages
RWGS reaction also requires water. The dry CO2/H2 system thus by product-derived surface species are established. While such
behaves as a autocatalyzed process. The figure also shows that ‘‘chromatographic’’ effects could contribute to the delays seen
the steady-state methanol rate is reduced as the partial pressure here, the following points support the ‘‘autocatalytic’’ effect of
of water in the feed is increased. The rate recovers when water is water suggested above. Most decisively, the two (supported and
removed, signifying a kinetic effect of water, not poisoning or sin- unsupported) catalysts have very similar numbers of copper sites.
tering. The optimum partial pressure of water required for meth- If filling of these surface sites by product-derived intermediates
anol synthesis rates under these conditions is evidently lower was a dominant cause of the observed delays, the supported cata-
than that used here. lyst should exhibit similar delay times to the unsupported catalyst
Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17 13

– even longer if the silica support surface area contributes to prod- 3.2. Water effects in CO/D2 mixtures
uct holdup. Secondly, when water is added to the reactants
(Fig. 1b–d), this water appears in the product stream with insuffi- Fig. 4 shows the results of similar initial reaction transients at
cient delay (assuming rapid adsorption) to populate a substantial 413 K using the supported Cu catalyst but with a 1:3 CO/D2 reac-
fraction of the surface. Furthermore, the surface adsorbate layer tant mixture instead of 1:3 CO2/D2. The upper panel (Fig. 4a) shows
is dominated by formate, readily made from the reactants (see that very little methanol is produced when exposed to a dry CO/D2
Fig. 3 and the following discussion), thus reducing the availability reactant, but when small amounts of D2O (mole fraction = 0.001)
of sites for water species. Finally, a maximum methanol production are added during this same experiment (panel b), methanol ap-
should occur if methanol product, held up during the induction pears and reaches a steady state of 1.6  10 4 lmol s 1 g 1. This
period, were displaced by water as it filled the surface. rate is about 60% lower than that eventually achieved in the
Fig. 3 shows the buildup of the IR intensity for Cu-bound surface CO2/H2 experiment in Fig. 2. The m/e = 18 signal from D2O is also
formate during the three experiments shown in Fig. 2. The inset at shown in these figures and show that the methanol responds
the lower right in panel b shows the ultimate (steady state) IR promptly to the appearance of the water in the reactor. This
spectrum of the O–C(D)–O stretch region. The bidentate formate suggests a more rapid formation of the key surface intermediates
symmetric stretch (b-DCOO/ms) is identified as the sharp feature that give methanol in moist CO/D2 environments than in CO2/D2.
at 1330 cm 1 and the asymmetric stretch feature at (b-DCOO/ma) The IR spectra taken at the end of each time sequence after
appears at 1590 cm 1[40]. The d C–H mode of the formate which reaching steady state are shown in Fig. 4c. No evidence for either
appears at 1360 cm 1 is red-shifted off scale with the D isotope the symmetric or asymmetric stretch of surface formate species
[40]. The IR spectra taken in the water co-feed experiments are is seen despite achieving a steady-state methanol synthesis rate
essentially identical to that measured in dry D2/CO2. The integrated approximately a factor of two lower than that achieved in CO2/H2.
intensity of the symmetric stretch band at 1330 cm 1 was used Assuming a linear response of the symmetric stretch IR intensity
to generate the intensity plots in Fig. 3. This intensity reaches a with coverage, the steady-state formate coverage at steady state
steady-state value estimated to correspond to a fractional coverage
of 0.20 [22]. In contrast to its effect on the methanol production
transient, the addition of water does not noticeably affect the
buildup of formate on the surface. The formate coverage is thus
decoupled from the methanol synthesis rate during these
(a)
experiments, a behavior more consistent with a spectator species. MS intensity
Our previous findings that bidentate formate is not directly hydro-
genated to form methanol is also consistent with this behavior
[36].
Copper-bound bidentate formate features dominate the IR spec-
m/e = 18
m/e = 34
tra under these conditions. Other than gas-phase features of carbon 0
monoxide, carbon dioxide and water, support-based features as-
signed to silanol (2755 cm 1), and adsorbed D2O (low intensity 0 400
broad band at 2580 cm 1) appear during these transients. These Time / s
features appear upon introduction of the reactant gas both with
and without added water in the reactant stream. In the latter case,
they presumably result from the water made in RWGS and meth-
(b)
MS intensity

anol synthesis.

m/e = 18
m/e = 34
0

500 800
Time / s
IR absorbance

(c)
0.01
IR intensity

without D2O

0
with D2O

1750 1650 1550 1450 1350 1250


Time / sec IR frequency (cm-1)

Fig. 3. IR intensity (integrated absorbance) of the symmetric stretching mode Fig. 4. MS intensities versus time during reaction of 10 sccm of D2/CO (3:1) at 6 bar
(1300–1380 cm 1 in the inset) versus time during methanol synthesis on a pre- over a CuSiO2 catalyst at 413 K (a) with no water added and (b) after adding
reduced silica-supported copper catalyst at 413 K. Reactant gas composition was 0.1 mol% D2O to the mixture at time = 500 s. The intensities of mass 34 (CD3O+) and
initially a D2/He mixture and then switched at the indicated time to a D2/CO2 18 (OD+) are both shown – the intensity scale for m/e = 18 is 1100 times greater
mixture (both with a 3:1 ratio at 6 bar) to which various amounts of water were than that for m/e = 34. (c) The IR spectra in the OCO formate stretch region under
added: (a) no water, (b) 0.1 mol%, (c) 0.2 mol% mixed. The inset IR spectrum was both conditions (i.e., at <400 s and >700 s, respectively). The same intensity scale
obtained during experiment c) after 500 s (i.e., steady state). (absorbance) is used as in the inset in Fig. 3.
14 Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17

in this experiment is <0.01, more than 20 times lower than that Table 1
in CO2/D2. When water is introduced to the D2/CO mixture, a small Methanol synthesis rates and isotopic composition of methanol made from CO2/13CO/D2
(1:1:6) reactant mixtures, 6 bar total pressure at the temperatures indicated.
amount of CO2 (10–20 ppm) is produced via water–gas shift. Such
a low concentration of CO2 is insufficient to produce measurable T/K Methanol Fraction 13C in Fraction 13C in
amounts of surface formate, reinforcing the general lack of TOF/10 6 s 1
methanol product exit CO2

correlation between methanol production and formate coverage. 433 5±2 0.87 ± 0.05 0.01 ± 0.001
It is theoretically possible that the methanol produced in the CO/ 453 21 ± 4 0.77 ± 0.06 0.01 ± 0.001
473 63 ± 12 0.63 ± 0.03 0.01 ± 0.001
D2 mixture arises from this small amount of CO2. The tracer 513 521 ± 40 0.22 ± 0.05 0.01 ± 0.001
experiments described in the next section address this point.
Other IR features include a broad absorption between 2500 and
2700 cm 1 (not shown), associated with OD on copper, and these in agreement with previous findings [29], but at the lower temper-
were observed to build in after the introduction of D2O. atures of this study, the preference shifts strongly from CO2 to CO
as the carbon source in methanol. As shown in Table 1, the amount
3.3. Competitive CO versus CO2 hydrogenation of scrambling of the isotope label between CO and CO2 was mini-
mal in these experiments. The raw data were corrected for minor
To address the origin of the methanol in the above experiments amounts of H–D exchange (<10% in control experiments with no
under these low temperature conditions, competitive tracer exper- carbon isotope labeling) with the walls of the mass spectrometer
iments were performed with 12CO2/;13CO/D2 mixtures. Steady- and possibly the catalyst support (see Experimental). Such
state carbon isotope tracer experiments have been previously used exchange inflates the apparent contribution by CO2 by converting
13
to separate the rates of CO versus CO2 hydrogenation under com- CD3OD (m/e 37) to 13CD3OH (m/e 36 and parent of 12CD3OD).
mercial conditions (500–550 K and 50 + bar total pressure) on The inclusion of additional water vapor (0.001 mol fraction) to
commercial Cu/ZnO catalysts [28–30]. These results showed that the reactant mixture after these steady-state values were achieved
CO2 is the preferred reactant. However, data at the lower temper- did not significantly alter the rates or the isotope distributions. The
atures of our study are not available nor was surface IR performed results in Table 1 are used to separate the contributions of CO and
simultaneously as done in the present study. As described in Sec- CO2 to the production of methanol. These are shown in Fig. 6 as
tion 2, our competitive experiments began with a period under turnover frequencies. The Arrhenius fits to the data yield the fol-
12
CO2/D2 (1:6) until a steady rate was achieved, whereupon 13CO lowing activation energies: 133 kJ mol 1 for CO2 and 66.3 kJ mol 1
was added. The water from reverse water–gas shift and methanol for CO.
synthesis reactions was thus already present when 13CO was intro- IR spectra taken during these experiments are shown in Fig. 7.
duced. Fig. 5 shows that after the introduction of 13CO at 433 K, the The O–C–O stretch region of the surface formate present in the
isotopic composition of the methanol product underwent a slug- D2/CO2 atmosphere at 433 K (spectrum a) was altered after CO
gish transient toward a new steady state, with the rate of methanol addition (spectrum b). Despite slight background irregularities in
production from the CO2 isotope being suppressed while methanol these spectra, it is clear that the O–(CD)–O symmetric stretch at
production from the CO isotope grew. The ultimate methanol pro- 1330 cm 1 appears to completely disappear upon addition of
13
duction rate in the final mixed reactant mixture was approxi- CO while a feature, while blue-shifted is retained at the fre-
mately ½ of that in the initial CO2/D2 (1:6) mixture. The length quency of the asymmetric stretch at 1600 cm 1. The low fre-
of time required to reach the new steady state was >2 h at 433 K quency absorption edge of the IR windows had shifted in the
and 30 min at 513 K. intervening time between the acquisition of the reference spec-
The fractional contents of 13C in the methanol product and CO2 trum and these experiments. In our previous paper [36], similar
in the reactor effluent at steady state for reaction at various tem- spectral features were assigned to monodentate formate in the
peratures are shown in Table 1. A clear trend is seen. At higher presence of co-adsorbates. Earlier work [44] on oxidized copper
temperatures, CO2 is the preferred source of carbon in methanol, documented similar changes, that is, (1) intensity changes in the
OCO bend and stretch modes consistent with reorientation of
the IR dipoles and (2) a non-significant (but measureable) shift of
the asymmetric feature from 1550 cm 1 to 1590 cm 1. This
spectrum is also consistent with that reported by Clarke and Bell
Methanol isotope fraction

1.0E-3

from CO2
Methanol TOF (1/s)

1.0E-4

from CO

1.0E-5

Time / sec

Fig. 5. Change of methanol carbon isotope fraction versus time after introduction of 1.0E-6
13
CO to a D2/CO2 reactant mixture. The intensity (I) fractions for 12CD3OD and 1.9 2 2.1 2.2 2.3 2.4
13
CD3OD are shown defined as the ratios I36/(I36 + I37) and I37/(I36 + I37), respec- 1000/T (K)
tively. The reactant mixture changed from D2/12CO2/He (6:1:1) to D2/12CO2/13CO
(6:1:1) at the time indicated. Cu/SiO2 catalyst at 433 K, 10 sccm total flow, 6 bar Fig. 6. Rates of methanol production as turnover frequencies (TOFs) from CO
total pressure. Intensities were corrected for H/D scrambling as described in Section (squares) and CO2 (circles) from the data in Table 1. Conditions: Cu/SiO2, 6 bar total
2.2. pressure, 13CO/CO2/D2 = 1:1:6, total flow 10 sccm.
Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17 15

alone or mixed, to methanol. In the case of methanol production


0.03 from CO at temperatures lower than 443 K, water is absolutely
necessary since very little methanol is made in dry D2/CO. The iso-
tope results above also show that CO can be directly reacted to
form methanol in competition with CO2 and in fact is the favored
reactant at these low temperatures. The methanol product in moist
D2/CO thus did not result from the small amounts of CO2 produced
IR absorbance

by water–gas shift in that experiment. Water or water-derived


intermediates may serve as reactants to form critical active inter-
mediate(s) or may serve as a ‘‘catalyst’’ for one or more hydrogena-
tion steps in both CO and CO2 hydrogenation. In the case of
0.0 methanol production from CO2, the induction time results in
Figs. 1 and 2 show that water must serve as a ‘‘catalyst’’ or reactant
for one or more CO2 hydrogenation steps. Furthermore, both RWGS
and methanol synthesis must be enabled by small amounts of
water since a prompt onset of RWGS would supply sufficient water
vapor to enable the methanol synthesis reaction. The parallel re-
verse water gas–shift reaction must be the major source of the
1650 1550 1450 1350 1250
water product at steady state, being 5–10 times faster than meth-
-1 anol synthesis under these conditions. Thus water is important for
IR frequency (cm )
the RWGS reaction as well. Indeed, if steady-state RWGS were
Fig. 7. Steady-state IR spectra during methanol synthesis at 433 K, 6 bar total reached rapidly in the experiments of Fig. 1, a ready supply of
pressure for the two reactant gas mixtures of Fig. 5: (a) D2/CO2/He (6:1:1) at water would have been available to develop methanol synthesis
time = 3000 s, (b) D2/CO2/13CO (6:1:1) at time = 8000 s. The intensity axis has on a similar time scale to the experiments with added water.
absorbance units. The two Cu-bound bidentate formate modes are labeled. Therefore, we believe that the RWGS reaction is likely also ‘‘auto-
catalyzed’’ by water or water-derived species. These results
strongly suggest a common intermediate for WGS and methanol
under similar conditions to ours [16], if their 1364 cm 1 feature for
synthesis and that water enables both pathways.
the H formate is assigned to the d C–H mode (with D substitution
Scheme 1 below encapsulates the mechanistic possibilities for
this feature shifts off-scale [40]). These are all consistent with our
methanol synthesis on copper from both CO and CO2, used here
observations in Fig. 7 and signify a bonding change resulting from
to put our results into perspective. The left hand side of the dia-
the presence of co-adsorbates in the presence of CO. Fig. 8 shows
gram shows the major steps in the formate pathway to methanol
that the change from bidentate to monodentate formate also oc-
from CO2 and the right-hand side shows a direct CO hydrogenation
curs over a relatively long timescale (1500 s), but shorter than
pathway to methanol via formyl. Although these general proposed
that required for the methanol isotope composition to reach the
mechanisms have been around for a long time, some new ideas
new steady state. Adjustments of the catalyst to new reaction con-
from recent DFT calculations described in Refs. [24,25] have been
ditions are sluggish at these temperatures and involve changes be-
incorporated here, including, for example, the breaking of one of
yond simple formate bonding/removal.
the C–O bonds in CO2 after the partial hydrogenation steps shown.
Across the top of the scheme is the mechanism of the water gas–
3.4. General discussion shift reaction (and its reverse) proposed based on DFT [31]. This
reaction interconverts CO and CO2. The proposed mechanism pro-
The results above show that water or water-derived surface ceeds through a carboxyl (H–O–C–O) intermediate in which the
species play a critical role in the conversion of both CO2 and CO, O–CO bond is activated. The formation of the carboxyl intermediate
from CO2 (via the step labeled A) or from CO (via the step labeled B)
are shown in the figure. A third methanol synthesis pathway which
involves this carboxyl intermediate has been proposed in our
recent paper [26] and is included as the middle column in the
scheme. This ‘‘middle way’’ for methanol formation arises as a
parasitic channel from this pool of water–gas shift carboxyl
IR integrated absorbance

intermediates. Earlier work, notably that of Clarke and Bell [16],


also supported the carboxyl intermediate for WGS. Rather than
have the carboxyl intermediate convert directly to methanol by
the hydrogenation steps in our paper, however, they proposed that
the carboxyl intermediate could isomerize to form the formate,
from which methanol was assumed to be produced.
This common carboxyl pathway, with formate as a spectator spe-
cies, has the following advantages in explaining a number of find-
ings in our data, as discussed next.
Water effect on CO2 hydrogenation: The DFT calculations in Ref.
[26] show that water can greatly assist the hydrogenation of CO2
to form carboxyl species, consistent with the observed autocata-
Time / sec lytic effect of water on CO2 reduction to both methanol and CO.
According to these calculations, the simple (dry) hydrogenation
Fig. 8. Evolution of the formate IR intensities (in absorbance units) with time upon
introduction of 13CO to D2 + 12CO2 for the same experiment as Fig. 6. Curve a = ma
of CO2 to form this intermediate is difficult (barrier = 1.25 eV)
(asymmetric OCO mode) integrated intensity 1480 cm 1 to 1650 cm 1. Curve b = ms and would be rate determining. The same calculations indicate that
(symmetric OCO mode) integrated intensity 1300 cm 1 to 1400 cm 1. water cannot assist the further conversion of formate by a further
16 Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17

Scheme 1. Mechanistic pathways for conversion of CO and CO2 to methanol over Cu. The water–gas shift mechanism through a carboxyl intermediate is seen across the top.
Pathways for the direct conversion of CO2 (through formate) and CO (through formyl) are shown in the left and right columns, respectively. The center column is the proposed
mechanism for methanol synthesis via the carboxyl intermediate from in Ref. [26]. See Section 3.4 for details.

C-hydrogenation step which is a difficult step in the formate mech- to the formation of the carboxyl from CO2/H2. The more rapid
anism. On the contrary, considering a wet copper surface for active achievement of catalytic steady state at low temperature for moist
reaction as most simply suggested by the results in Fig. 1, calcula- CO/H2 reactant than moist CO2/H2 is consistent with this observa-
tions showed that the presence of water/hydroxyl increased the tion. Since the observed activation energies arise from a complex
formate hydrogenation barrier from 1.2 eV to 1.8 eV. Including combination of temperature dependent coverages and reaction
the effects of water, the carboxyl pathway has an overall barrier barriers, a full microkinetic description is required to understand
for methanol synthesis of 1.08 eV versus 1.8 eV for the formate the origin of this behavior.
pathway. The fact that the RWGS reaction is also assisted by water Similar rates of methanol synthesis in CO/H2 and CO2/H2 in the
is a natural consequence of a common intermediate, whereas other presence of water: A common intermediate accessed by both CO
mechanisms require separate explanations for the two reaction and CO2, combined with the relatively similar ease of formation
channels. We note that the mechanism in Ref. [34] includes from either reactant, provides a simple explanation for the similar
reactions between carboxyl and CO2 + either OH or H2O in their rates of methanol in the two syngas compositions. Admittedly, this
water–gas shift mechanism, that is, reactions 16 and 17 in is not a strong argument, given the very different surface adlayer
Table 1, Ref. [34]. compositions under the two conditions.
Water effect on CO hydrogenation: In order to explain our results, Taken altogether, the common carboxyl intermediate mecha-
a water-assisted step or strong coadsorbate effect would be nism naturally correlates the results. This hypothesis is strength-
required in one of the steps of this mechanism. However, water ened by the DFT calculations in Ref. [26] which show both the
is not directly involved in the traditional direct CO hydrogenation energetic feasibility of this pathway and also identifies an impor-
mechanism. Instead, OH is required to form the carboxyl interme- tant role for water-derived species.
diate from CO according to theoretical investigations of the WGS All of our results could of course have alternate explanations
reaction [31] and, thus, the water effect on methanol production with separate mechanistic channels for CO and CO2 conversion to
from the common intermediate is naturally explained. methanol. In doing so, however, the effects of water on both chan-
12
CO2–13CO–D2 isotope tracer experiments: The simultaneous nels, now largely ignored, would need to be included. Furthermore,
participation of both CO and CO2 in methanol synthesis is a natural difficulties still remain in the formate pathway for CO2 conversion
consequence of a common intermediate in the faster WGS reaction. to methanol, notably (1) our DFT calculations failed to find a water-
In this circumstance, the common intermediate will be predomi- assisted step in the formate conversion pathway to methanol, and
nately produced from the more facile source of the intermediate, (2) methanol was not produced in ‘‘titration’’ experiments where
that is, CO or CO2, under the reaction conditions. For instance, if direct, dry, hydrogenation of formate adlayers on copper was at-
step A in Scheme 1 were much faster than step B, the common tempted [36]. More complete experimentation and microkinetic
intermediate would exchange carbon more rapidly with the CO2 modeling are required to sort all this out. From the DFT energies
reactant than with CO. In this case the isotopic composition of in [25,26] and in modeling by Mavrikakis and coworkers, the stea-
the intermediate, and the methanol produced from it, would be dy-state coverages by the carboxyl intermediates are predicted to
dominated by that of CO2. Furthermore, step B would also be the be very low and are therefore unlikely to ever by detected in
slow step in the (reverse) water–gas shift reaction in this instance. in situ measurements. This would explain the invisibility of such
CO would likewise dominate if step B were relatively easier. adsorbed intermediates in the operand FTIR data for either
According to this model, the results in Table 1 indicate that car- CO/D2/H2O or CO2/D2/H2O mixtures presented here and elsewhere.
boxyl formation from CO becomes easier than its formation from Other co-adsorbates, including weakly bound ones such as
CO2 as temperature is lowered, that is, the effective activation en- water, could also be playing roles not yet appreciated. Water has
ergy for CO2 activation is higher than that for CO activation. This in been shown to assist in support-bound formate decomposition to
turn indicates that at low temperatures, the difficult process in the both CO and CO2 [45]. Facile formate-carboxyl interconversion, as
RWGS mechanism shifts from converting the carboxyl to CO + H2O suggested by Clarke and Bell [16], would provide new pathways
Y. Yang et al. / Journal of Catalysis 298 (2013) 10–17 17

for formate to contribute to methanol synthesis as a reservoir spe- [5] D.R. Palo, R.A. Dagle, Holladay, Chem. Rev. 107 (2007) 3992.
[6] C.S. Song, Catal. Today 115 (2006) 2.
cies even if carboxyl proves to be the key reactant for methanol
[7] G.A. Olah, A. Goeppert, G.K.S. Prakash, J. Org. Chem. 74 (2009) 487–498.
synthesis. Such steps could be also assisted by coadsorbate effects, [8] C.T. Campbell, G. Ertl, H. Kuipers, J. Segner, J. Chem. Phys. 73 (1980) 5362.
and future DFT studies should shed more light on such effects. [9] C.T. Campbell, K.A. Daube, J. Catal. 104 (1984) 109.
[10] M.E. Domagala, C.T. Campbell, Catal. Lett. 9 (1991) 65.
[11] K.H. Ernst, C.T. Campbell, G. Moretti, J. Catal. 134 (1992) 64.
4. Conclusions [12] C.V. Ovsen, P. Stoltze, J.K. Norskov, C.T. Campbell, J. Catal. 134 (1992) 445.
[13] I. Chorkendorff, P.A. Taylor, P.B. Rasmussen, J. Vac. Technol. A 10 (1992) 2277.
[14] P.B. Rasmussen, P.M. Holmblad, T. Askgaard, C.V. Ovesen, P. Stoltze, J.K.
(1) A common intermediate is implicated in methanol forma- Norskov, I. Chorkendorff, Catal. Lett. 26 (1994) 373.
tion from CO and CO2. [15] P.B. Rasmussen, M. Kazuta, I. Chorkendorff, Surf. Sci. 318 (1994) 267.
(2) Carboxyl is a strong candidate for this intermediate. [16] D.B. Clarke, A. Bell, J. Catal. 154 (1995) 314.
[17] T.S. Askgaard, J.K. Nørskov, C.V. Ovesen, P. Stolze, J. Catal. 156 (1995) 229.
(3) Water and water-derived species are critical reactants and
[18] P.A. Taylor, P.B. Rasmussen, I. Chorkendorff, J. Chem. Soc. Faraday Trans. 91
co-reactants in both WGS and methanol synthesis on (1995) 1267.
copper. [19] J. Yoshihara, S.C. Parker, A. Schafer, C.T. Campbell, Catal. Lett. 31 (1995) 313.
[20] J. Yoshihara, C.T. Campbell, J. Catal. 161 (1996) 776.
(4) The preferred carbon source for the reaction changes from
[21] B. Sakakini, J. Tabatabaei, M.J. Watson, K.C. Waugh, F.W. Zemicael, Farad. Disc.
CO2 to CO as temperature is lowered. 105 (1996) 369–376.
[22] T. Fujitani, I. Nakamura, T. Uchijima, J. Nakamura, Surf. Sci. 383 (1997) 285.
[23] J. Szanyi, D.W. Goodman, Catal. Lett. 10 (1991) 383.
[24] Y.X. Yang, J. Evans, J.A. Rodriguez, M.G. White, P. Liu, Phys. Chem. Chem. Phys.
Acknowledgments 12 (2010) 9909.
[25] L.C. Grabow, M. Mavrikakis, ACS Catal. 1 (2011) 365.
This study was performed at the Institute for Integrated Catalysis [26] Y. Zhao, Y. Yang, C. Mims, C.H.F. Peden, J. Li, D.H. Mei, J. Catal. 281 (2011) 199.
[27] D.H. Mei, L. Xu, G. Henkelman, J. Catal. 258 (1) (2008) 44–51.
(IIC) at Pacific Northwest National Laboratory (PNNL), and partially [28] Yu B. Kagan, L.G. Liberov, E.V. Slivinsky, S.M. Lockev, G.I. Lin, A.Ya. Rosovsky,
funded by a Laboratory Directed Research and Development (LDRD) A.N. Bashkirov, Dokl. Akad. Nauk. SSSR 222 (1975) 1093.
grant as part of the Catalysis Initiative program administered by [29] G.C. Chinchen, P.J. Denny, D.G. Parker, M.S. Spencer, D.A. Whan, Appl. Catal. 30
(1987) 333.
PNNL. The work was carried out in the Environmental Molecular [30] V.E. Ostrovskii, Catal. Today 77 (2002) 141.
Sciences Laboratory (EMSL) at PNNL, a National Scientific User facil- [31] A.A. Gokhale, J.A. Dumesic, M. Mavrikakis, J. Am. Chem. Soc. 130 (2008) 1402.
ity supported by the US Department of Energy Office of Biological [32] C-H Lin, C-L. Chen, J.H. Wang, J. Phys. Chem. C 115 (2011) 18582–18588.
[33] R. Burch, A. Goguet, F.C. Meunier, Appl. Cat. A: General 409–410 (2011) 3–12.
and Environmental Research. PNNL is operated by Battelle for the
[34] R.J. Madon, D. Braden, S. Kandoi, P. Nagel, M. Mavrikakis, J.A. Dumesic, J. Catal.
US Department of Energy. C.T.C. (Grant number DE-FG02- 281 (2011) 1–11.
96ER14630) and C.H.F.P. would like to acknowledge the Depart- [35] J. Nakamura, J.M. Campbell, C.T. Campbell, Faraday Trans. 86 (1990) 2725–
2734.
ment of Energy, Office of Basic Energy Sciences, Division of Chemical
[36] Y. Yang, C.A. Mims, R.S. Disselkamp, Ja-Hun Kwak, C.H.F. Peden, C.T. Campbell,
Sciences, Biosciences and Geosciences, for support of their partici- J. Phys. Chem. C 114 (2010) 17205.
pation in this work. C.A.M. gratefully acknowledges PNNL support [37] R. Bardet, J. Thivolle-Cazat, Y. Trambouze, C.R. Hebd, Seances Acad. Sci. Ser. C
for his participation in the IIC as a visiting professor. Helpful discus- 299 (1984) 423.
[38] K. Klier, V. Chatikavanij, R.G. Herman, G.W. Simmons, J. Catal. 74 (1982) 343.
sions with Enrique Iglesia are gratefully acknowledged. [39] Y. Yang, R.S. Disselkamp, C.T. Campbell, J. Szanyi, C.H.F. Peden, J.G. Goodwin Jr.,
Rev. Sci. Instrum. 77 (2006) 94–104.
References [40] Y. Yang, C.A. Mims, R.S. Disselkamp, C.H.F. Peden, C.T. Campbell, Top. Catal. 52
(2009) 1440.
[41] G.C. Chinchen, C.M. Hay, H.D. Vandervell, K.C. Waugh, J. Catal. 103 (1987).
[1] G.C. Chinchen, P.J. Denny, J.R. Jennings, M.S. Spencer, K.C. Waugh, Appl. Catal.
[42] M.-J. Luys, P.H. van Oeffelt, P. Pieters, R. Ter Veen, Catal. Today 10 (1991) 283.
36 (1988) 1–65.
[43] M.J. Luys, P.H. van Oeffelt, W.G.J. Brouwer, A.P. Pijpers, J.J.F. Scholten, Appl.
[2] X.M. Liu, G.Q. Lu, Z.F. Yan, J. Beltramini, Ind. Eng. Chem. Res. 42 (2003) 6518.
Catal. 46 (1989) 161.
[3] R. Farrauto, S. Hwang, L. Shore, W. Ruettinger, T. Giroux, Y. Liu, O. Ilinich, Ann.
[44] G.J. Millar, C.H. Rochester, K.C. Waugh, J. Chem. Soc. Faraday Trans. 87 (1991)
Rev. Mater Res. 33 (2003) 1 (General review of MeOH synthesis).
2795.
[4] J. Ma, N.N. Sun, X.L. Zhang, N. Zhao, F.K. Mao, W. Wei, Y.H. Sun, Catal. Today.
[45] T. Shido, A. Yamaguchi, K. Asakura, Y. Iwasawa, J. Mol. Catal. 163 (2000) 67.
148 (2009) 221.

You might also like