Вы находитесь на странице: 1из 9

Fuel Processing Technology 146 (2016) 3947

Contents lists available at ScienceDirect

Fuel Processing Technology


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

Preparation of Cu/ZnO-based heterogeneous catalysts by photochemical


deposition, their characterisation and application for methanol synthesis
from carbon dioxide and hydrogen
M.B. Pori a, B. Likozar a, M. Marinek b, Z. Crnjak Orel a,
a
b

National Institute of Chemistry, Hajdrihova 19, SI-1000 Ljubljana, Slovenia


Faculty of Chemistry and Chemical Technology, Vena pot 113, SI-1000 Ljubljana, Slovenia

a r t i c l e

i n f o

Article history:
Received 7 October 2015
Received in revised form 11 February 2016
Accepted 15 February 2016
Available online xxxx
Keywords:
Photochemical catalyst synthesis
Heterogeneous copper catalysts
Structureactivity correlation
Carbon dioxide activation
Methanol synthesis reaction

a b s t r a c t
Binary Cu/ZnO methanol synthesis catalysts were prepared by photochemical method, whereby CuO crystallites
were chemically anchored on the surface of ZnO support using UV irradiation, and the prepared CuO/ZnO
composites were subsequently reduced. The amount of CuO crystallites was varied with the concentration of
Cu2+ ions used and the time of UV irradiation in order to obtain a wide range of Cu crystallite size. The activity
of the catalysts was evaluated for the methanol synthesis reactions from CO2 and H2 at mild reaction conditions
(250 C; 1 bar), and was compared as a relative activity to the commercial methanol synthesis catalyst (Cu/ZnO/
Al2O3). The activity of the prepared catalysts was up to 18-fold higher in comparison with the commercial ones. A
superior activity was disclosed in terms of the abundant Cu and ZnO surfaces, exposed to reacting molecules. The
size of Cu crystallites greatly inuenced the formation of methanol, which is favored over the catalysts with
smaller Cu crystallites. The latter is proposed to arise from the high dispersion of Cu crystallites on ZnO support,
and thus from the enhanced contact between Cu and ZnO particles. The prepared catalysts also showed a high
long-term stability due to the small amount of water and carbon monoxide, produced as a by-product.
2016 Elsevier B.V. All rights reserved.

1. Introduction
With rapid changes and the development in industry, the energetic
and environmental aspects have become two major concerns of several
processes, including the conventional oil and gas rening, cement production, etc. It is well known, that CO2 is one of the most widespread
greenhouse gases, pertinent to global warming issues. For this reason,
great efforts have been invested to converting the emitted CO2 to useful
chemicals and fuels [13].
One of the most interesting ways of the mentioned CO2 emission'
reuse would be their conversion by methanol synthesis. Considering
this, methanol poses itself as one of the most appealing platform
chemicals for further use in ne chemicals' syntheses, as a liquid energy
carrier for direct- or reforming-coupled fuel cells, and for several further
exploitation routes [4]. When methanol is used as a fuel, it may be considered rather environment-friendly compared to several other sources
of energy (foremost coal and oil); besides, due to its liquid form, it can
provide a convenient fuel source in fuel cell application, as well as in
conventional energy storage. Its versatile nature allows for further
transformation into formaldehyde (and the application in resin

Corresponding author.
E-mail address: zorica.crnjak.orel@ki.si (Z. Crnjak Orel).

http://dx.doi.org/10.1016/j.fuproc.2016.02.021
0378-3820/ 2016 Elsevier B.V. All rights reserved.

synthesis), dimethyl ether, or to be used as a solvent in several commercial processes) [4].


Commercially, methanol is produced from syngas, containing CO,
CO2 and H2, over a ternary Cu/ZnO/Al2O3 catalysts, prepared with the
co-precipitation method with sodium carbonate. The process is executed under high pressures (a lower limit of about 510 MPa) and high
temperatures (493573 K) [2]. Catalyst synthesis route is based on the
preparation of Cu and Zn (carbonate) precursors by co-precipitation,
using metal nitrate solutions and sodium carbonate as precipitation
agent. The precipitate is aged, recovered and thermally decomposed in
order to obtain a ne mixture of oxides. After the reduction of CuO,
the catalyst has a typical atomic Cu/Zn ratio near 70:30, whereby the
molar Al content is 1020% [5]. The conventional co-precipitation method, no matter how widespread, has some disadvantages. Additionally,
the conventional catalysts are tailored for syngas conversion, the latter
containing comparably low CO2 contents. Recently, the methanol synthesis by CO2 hydrogenation (overall reaction: CO2 + 3H2
CH3OH + H2O) has received a worldwide research interest, since it is
an effective way of CO2 activation and conversion [6]. However, when
the conventional Cu-based catalyst is used for CO2 hydrogenation to
methanol, the yield of methanol is much lower than that obtained
from syngas under similar reaction conditions [2]. The reasons can be
traced back to the deactivation of the Cu/ZnO/Al2O3 catalysts due to
the production of water as a by-product, which decreases the yield of

40

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

methanol production dramatically [7,8]. Besides water, the main important mechanism of Cu-based catalysts deactivation is sintering, related
to the low melting temperature of Cu crystallites [9,10].
Many research groups have adopted unconventional preparation
methods to improve the activity, selectivity, poison resistance and lifetime of the CuZnO-based catalytic systems. It is common to all these
studies that a satisfactory catalyst performance may be achieved by a
high dispersion of the active phase over support that is by the interface
contact between Cu and ZnO. For example, Karelovic et al. [1] found out
that the catalysts, prepared by the citrate method, feature a high methanol selectivity, which arises from the high inter-dispersion between Cu
and ZnO. Wang et al. [2] analogously prepared a series of Cu/ZnO/Al2O3
catalysts through the decomposition of M(Cu,Zn)ammonia complexes
under sub-atmospheric pressure to enhance the activity, whereas Jingfa
et al. [11] emphasized a superior activity of oxalate-co-precipitated catalyst materials. All these works inherently aim at Cu/ZnO interface area
increase.
In the present study, a new preparation method for Cu/ZnO catalysts,
based on the photochemical deposition procedure of CuO on the previously prepared ZnO substrate, and subsequent reduction, is described.
The exhibited activity and stability toward CO2 hydrogenation to methanol of the prepared catalysts is investigated, and additionally, some
fundamental structural factors controlling the stability and activity of
the employed CuZnO systems are described.

2. Experimental
2.1. Preparation of catalysts
Zinc oxide (ZnO) particles, which were used as the substrate material for the photochemical deposition of CuO in this work, were synthesized by the co-precipitation method, described by Bitenc et al. [12].
The initial concentrations of Zn2+ and urea were 0.01 M and 0.05 M, respectively, while experiments were carried out in laboratory bottles,
utilizing the working volume of reaction mixtures of 0.8 L. Experiments
were performed in an oven, preheated at 90 C for 4 h. The resulting
solid was ltered off, washed with demineralized water and dried overnight (16 h) at 60 C. The samples were subsequently thermally treated
at 300 C for 2 h in air (calcination process was optimized in previous
work by Bitenc et al. [12]). The prepared ZnO particles were used as
the substrate material for the photochemical deposition of copper
oxide (CuO).
In a typical photochemical coating process of CuO, 25 mg of the prior
synthesized ZnO was dispersed in 50 mL of the demineralized water
using the ultrasound for 10 min. To such dispersion, the aqueous solution
of copper nitrate was admixed (prepared from Cu(NO3)2 2.5H2O
(98 wt.% purity, Sigma-Aldrich) to obtain 0.501.00 mM concentration
of Cu2+, respectively, and subsequently irradiated using UV source
(monochromatic; 365 nm wavelength). Photochemical process was
maintained at the room temperature (23 1 C), while irradiation time
varied from 1.5 h to 12.0 h. The resulting solid was ltered off, rinsed
with demineralized water and yet again dried overnight (16 h) at 60 C.
The detailed experimental conditions for obtaining the resulting catalytic
materials, denoted AE, correspondingly, are given in Table 1.

Table 1
Experimental conditions for photochemical deposition of CuO on ZnO nano-sheets.
Sample
index

Batch volume
(mL)

Deposition
time (h)

Cu2+ concentration
(mM)

A
B
C
D
E

250
250
250
250
250

1.5
1.5
1.5
4.0
12.0

0.50
0.75
1.00
1.00
1.00

2.2. Characterization of catalysts


The synthesized products were characterized by eld-emission
scanning electron microscopy (FE-SEM; Zeiss Supra 35 VP), equipped
with the energy-dispersive X-ray (EDX) analyzer for composition determination. X-ray diffraction analyses (XRD) were carried out in order to
determine the crystallinity and crystallite size of the samples using Siemens D-500 X-ray diffractometer PANalytical X'Pert PRO (employing
the CuK1 radiation with = 1.5406 ). Fourier transform infrared
(FTIR) spectra were obtained using a FTIR spectrometer (PerkinElmer
2000) in the spectral range between 350 and 700 cm1 with the spectral resolution of 2 cm1 (samples were prepared as KBr pellets in N2
atmosphere). BET specic surface area was determined using
Micromeritics Gemini II 2370 Surface Area Analyzer according to the
BET multipoint method with ultrahigh purity nitrogen (5.5) being
used as the adsorbing gas.
2.3. Application of catalysts for methanol synthesis from CO2 and H2
The prepared powder mixtures (0.010.3 g) were consequently subjected to temperature-programed reduction (TPR) characterization.
Samples were rstly heated from 25 to 120 C in Ar (5.0) ow of
50 mL/min, using the heating rate of 10 K/min, and maintained at the
nal temperature for 60 min. Afterwards, they were cooled to 15 C
using identical cooling rate, atmosphere and ow rate, and maintained
at this temperature for 5 min. Subsequently, they were re-heated from
15 to 250 C in the 4.9 vol.% H2/Ar (5.0) ow of 50 mL/min, using
the heating rate of 1 K/min, and maintained at the nal temperature
for 180 min, and lastly, cooled to 25 C using 20 mL/min Ar ow rate.
Micromeritics AutoChem II Chemisorption Analyzer apparatus was
used for TPR, employing thermal conductivity detector (TCD) to determine the H2 consumption during reduction.
Following the reduction, the samples were subjected to model reaction system using the aforementioned apparatus. Firstly, they were
heated from 25 C to 250 C using a heating rate of 5 K/min and the constant H2 (5.0) ow rate (51 mL/min), after which they were maintained
at the nal temperature for 505 min; that is 5 min in pure H2 and then
using additional 100 dosing cycles (5 min for each dosing cycle), where
CO2 was added to H2 (the CO2 ow rate of 17 mL/min with each dosing
volume of 0.5324 mL). Mass spectrometry (MS) (ThermoStar, Pfeiffer
Vacuum) was used to determine the amount of the reactants and
products in efuent gases (maintained at 130 C) by monitoring the
characteristic m/z signals for CH4 (m/z = 15.00), H2O (m/z = 18.00),
CO (m/z = 28.01), formaldehyde (m/z = 30.00), methanol (m/z =
32.04), CO2 (m/z = 44.00) and dimethyl ether (m/z = 46.07). Prior to
further calculations, the m/z signals were normalized by subtracting
the MS response of an empty quartz reactor, subjected to the same experimental conditions as synthesized catalytic materials. Catalytic activity and selectivity in terms of the formation/consumption of these
species were evaluated at the beginning, middle and end the mentioned
100 dosing cycles, examining the temporal stability of activity as well.
Activity of the prepared catalysts was compared to the activity of commercial catalyst (Cu/ZnO/Al2O3; HiFUEL W230, Alfa Aesar; grinded) and
expressed as a relative activity ratio. After reactions, the samples were in
turn cooled in He (6.0) atmosphere to 25 C using the ow rate of
30 mL/min.
3. Results and discussion
3.1. Synthesis and characterization of Cu/ZnO composites
3.1.1. Composite structure characterization
Fig. 1 shows the SEM image of ZnO particles prepared with the
above-described homogeneous precipitation method. Such particles
were used as the substrate material for the subsequent photochemical
deposition of the CuO nanoparticles (NP). It is evident that ZnO particles

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

41

Fig. 1. FE-SEM micrographs of ZnO particles (a), and FE-SEM micrographs of CuO/ZnO composites prepared at various durations of UV irradiation and concentrations of Cu2+ ions that is
micrographs of sample A (b), B (c), C (d), D (e) and E (f).

are composed of nano-sheets with the thickness around 30 nm. In the


work of Bitenc et al. [12] it was established that the surface area of
ZnO particles prepared by homogeneous precipitation is relatively
high due to the nano-porous structures of individual sheets, which is
formed during the thermal treatment of zinc hydroxide carbonate
(ZHC) at 300 C for 2 h. Nano-porous sheets result in a relatively large
specic surface area, which was determined as 44 m2/g for the prepared
ZnO substrate. A large surface area of ZnO particles is a highly desired
property, since catalytic activity among others depends on the specic
surface area of a support [13].
Morphological characteristics of the prepared samples are somewhat changed after the photochemical precipitation process of CuO
(Fig. 1bf). Evidently, CuO deposits most intensely on the edges and
on the outskirts of individual ZnO nano-sheets. Such distribution is
caused by at least two effects; the fact that i) CuO is formed through
the irradiation process, which is hindered in shady regions, and ii) a relatively narrow spacing between the nano-sheets of ZnO aggravates the
transport of Cu2+ ions within the regions between the individual sheets
of ZnO. The effect of the various concentrations of Cu2+ ions in the precursor solution on the growth of the CuO/ZnO composites under the
xed irradiation time of 1.5 h is demonstrated in Fig. 1b, c and d. With
the increasing concentration of Cu2+ ions there is an increment of the
deposited amount of CuO nanostructures. When the initial concentration of Cu2+ ions is 0.5 mM/L (Fig. 1b), CuO nanoparticles attach uniformly on the edges of ZnO nano-sheets. With the increasing initial
concentration of Cu2 + ions (CCu2 + = 0.75 mM/L), CuO NP started to

grow also inside ZnO nano-sheets (Fig. 1c), whereas the outer surface
of ZnO nano-sheets is completely covered with CuO NP when the initial
concentration of Cu2+ reached 1.00 mM (Fig. 1d). Fig. 1df shows the
images of CuO/ZnO composites prepared using various irradiation
times of UV light, respectively. When the irradiation time was increased,
more CuO NP decorated the surface of ZnO nano-sheets, meaning that
after the 4 h and 12 h of irradiation process, CuO NP are densely packed
on the surface of ZnO nano-sheets (Fig. 1e and f).
The typical X-ray diffraction patterns of the CuO/ZnO composites
and Cu/ZnO composites are shown in Figs. 2 and 3. CuO/ZnO composites
are characterized by the presence of two well-crystallized phases
(Fig. 2); specically, CuO (JCPDS 00-005-0661) and ZnO (JCPDS 00005-0664). No other crystalline phases were detected. The intensity of
the corresponding diffractions for CuO grows with the increasing atomic
ratio of Cu/Zn. For the CuO/ZnO composites with the Cu/Zn atomic ratio
in the range of 0.30.4, CuO diffraction peaks are hardly visible, indicating that the CuO NP in these catalysts should have a comparatively small
crystallite size. By contrast, larger CuO crystallite sizes are obtained for
the catalysts with the Cu/Zn atomic ratio of 0.74 and 0.91, since the corresponding signals are slightly more intense and narrower, suggesting a
higher tendency toward the formation of larger CuO clusters with the
raise in Cu loading.
The X-ray diffraction patterns of the Cu/ZnO composites after
methanol synthesis are shown in Fig. 3. Analyses revealed that Cu is
mostly present in the form of Cu0 (JCPDS 00-003-1015) with the
exception of composites A and D, where a minor amount of CuO species

42

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947


Table 2
Physical properties of CuO/ZnO and Cu/ZnO composites.
Catalyst

A
B
C
D
E
ZnO

CuO/ZnO before reduction

Cu/ZnO after
reduction

Cu/ZnO after
reaction

BET
(m2/g)

d(CuO)a
(nm)

WCu/WZnb
(/)

d(Cu)a
(nm)

d(Cu)a
(nm)

WCu/WZnb
(/)

58.3
57.0
55.7
53.8
52.6
44.0

29.7
32.3

0.30
0.35
0.40
0.74
0.91

43.3
46.6
55.1
56.4
62.4

55.4
59.8
77.1
89.5
91.6

0.24
0.36
0.42
0.61
0.96

a
Crystallite size of CuO and Cu was calculated from full width at half maximum
(FWHM) of XRD diffraction peaks at 38.9 and 43.4, respectively, using the Scherrer
equation.
b
Atomic ratio.

Fig. 2. XRD spectra of CuO/ZnO composites.

was detected. The source of oxygen for Cu0 oxidation (e.g.


Cu + H2O CuO + H2) is H20, which is formed during methanol synthesis reactions as a by-product [14]. While transient oxidation reactions may take place, CuO is slowly reduced back to its initial state by
H2 and CO [15], explaining the absence of CuO in composites B, C and E.
The size of the CuO and Cu crystallites after reduction and methanol
synthesis reactions was calculated using the Scherrer equation, d = K /
(B cos()), in which K is the shape factor with the value 0.9, B the structural broadening, and represents the most intense X-ray diffraction
peak of a sample [3]. The size of the CuO NP for samples D and E was determined at the diffraction angle of 38.9, whereas it could not have
been evaluated for samples A, B and C due to the poor crystallinity of
CuO NP. The crystallite size of Cu NP was thus established at the diffraction angle of 43.4. According to the results gathered in Table 2, the size
of Cu crystallites, obtained after methanol synthesis reactions, is larger
in comparison to Cu crystallites prior to it. The latter could be explained
through the sintering of Cu particles during methanol synthesis, which
is related to a rather low Htting and Tammann temperatures of Cu,
i.e. the mobility of the Cu atoms on catalysts surface becoming signicant [16].

Fig. 3. XRD spectra of Cu/ZnO composites after reduction and methanol synthesis reaction.

The FTIR spectra of CuO/ZnO composites were carried out in the


range of 700370 cm1 (Fig. 4). The presence of ZnO and CuO in the
mentioned products is conrmed by the bands at 418 cm 1 and
520 cm1, respectively, as well as by a pronounced shoulder, characteristic for CuO at 590 cm1 [12,17]. The intensity of bands, assigned to
CuO, increases with the elevated concentration of Cu2+ ions in precursor solution, used for photochemical process, and the time of irradiation.
These results conrmed that the amount of the deposited CuO NP on
ZnO nano-sheets is indeed increased with both the time of irradiation
and the initial concentration of Cu2+.
3.1.2. Surface area characterization
According to the results in Table 2, any CuO/ZnO composite exhibits
a higher value of BET surface area in comparison to neat ZnO, and thus
the deposition of the CuO NP on ZnO nano-sheets apparently increases
specic surface area. In order to investigate the correlation between the
amount of the deposited CuO and BET surface area, EDX (point) analyses were performed. It appears that BET surface area approximately inversely depends on the amount of the deposited CuO (Table 2). The
latter trend could be explained with a partial blockage of porous ZnO
surface by loading CuO species. Nonetheless, given the specic surface
area of neat ZnO, a maximum in BET surface apparently exists, where
an optimal trade-off between surface area decrease due to pore blocking
and its increase by rendering the same surface rougher owing to CuO
deposition is reached. Our results are consistent with the ones, obtained
in the work of Agarwal et al. [18]. In this study, the BET specic surface

Fig. 4. FTIR spectra of CuO/ZnO composites prepared with various concentrations of Cu2+
ions and times of irradiation.

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

of CuO/ZnO/Al2O3 composites drops with the increasing CuO loading on


substrate surface.
3.1.3. Composite reduction characterization
In order to establish the reduction behavior of the prepared catalysts
and to determine the proper reduction protocol for testing, TPR measurements were performed by monitoring H2 consumption (Fig. 5).
The relative volumes of the H2 consumed per the mass of catalysts
(VH2) are shown in Table 3 and are dened as the areas under the TPR
curves in range of 50250 C. The amount of the consumed H2
corresponded to the reduction of Cu2 + in all the catalysts, prepared
using photochemical deposition method. Additionally, onset (To) and
maximum (Tmax) temperatures are given in Table 3. As shown in
Fig. 5, all of the systems with different Cu/Zn atomic ratios exhibited a
single H2 consumption peak in the 100160 C region, and showed a
maximum around 140 C. The presence of only one reduction peak
was in accordance with the literature data, although the prepared catalysts were reduced at much lower temperatures when compared to a
standard bulk CuO (297 C) [19]. The narrow and almost symmetrical
H2 consumption peaks, located at lower temperatures, were ascribed
to the reduction of the well-dispersed minute CuO crystallites with a
narrow size distribution [19,20].
As observed in Fig. 5, the reduction behavior of the synthesized catalysts strongly depended on Cu/Zn atomic ratio. An increase of the volume of H2, consumed for reduction was observed paralleling the ratio.
The reduction of the catalysts with the Cu/Zn ratio of 0.91 required a relatively high specic volume of H2 (19.4 a.u. g1). Conversely, signicantly less H2 was consumed for the reduction of the materials with
Cu/Zn ratio decreasing to the value of 0.74. Thus, the catalysts with a
low Cu/Zn ratio of 0.3 required the lowest amount of H2 (4.6 a.u. g1).
The trend observed could be explained with the augmentation of the
CuO crystallites with increasing Cu/Zn atomic ratio. Larger particles
can overlap with each other, making reduction more difcult due to a
lower relative surface area exposed to H2 [21]. Furthermore, a decrease
in Cu/Zn atomic ratio paralleled a straightforward elevation in To
(Table 3). The latter suggested that the formation of nucleating metal
centers (reected through To) was more facile for larger CuO crystallites,
which were characterized by having only a slight interaction with ZnO.
The dispersed CuO particles, which are in a good contact with ZnO, shift
To towards higher respective values.
3.2. Catalytic activity of Cu/ZnO composites
Catalytic activity tests were performed with neat ZnO and the prepared CuO/ZnO dispersions. According to the results, obtained by

Fig. 5. Temperature-programed reduction proles of prepared catalysts.

43

Table 3
Summary of TPR data for catalysts, prepared with photochemical method.
Catalysts

To (C)a

Tmax (C)a

VH2 (a.u. g1)a

Cu0.30Zn1
Cu0.35Zn1
Cu0.40Zn1
Cu0.74Zn1
Cu0.91Zn1

115
108
117
106
100

140.0
145.0
145.0
139.5
140.6

4.6
5.9
6.7
9.9
19.4

a
Onset temperature of reduction (To), temperature of peak maxima (Tmax) and relative
volume of consumed hydrogen (VH2).

activity tests, neat ZnO is basically inactive toward the synthesis of


methanol, while the rate of the methanol formation increases signicantly when the dispersion of two oxides, CuO and ZnO, is applied.
These results show that a sufcient contact between both phases is
among others one of the key factors responsible for tailoring a satisfactory methanol synthesis activity, as already stressed in some earlier
studies [1].
3.2.1. Inuence of Cu crystallite content
The study of the inuence of the Cu content on catalytic activity was
conducted over the materials, prepared with photochemical method
(Fig. 6). As expected, with (notably) increasing the amount of Cu in
the prepared catalysts, the methanol-based activity, calculated at the
beginning of synthesis reaction, decreased.
For the catalyst B with 6.1 wt.% Cu a maximum relative activity for
methanol formation (17.5) was observed. Upon increasing Cu loading
(up to 17.4 wt.%), the activity for methanol formation dropped down
to the value of 8.8. In the work of Mehta et al. [22], the authors claimed
that in the catalysts with a low Cu content, Cu crystallites were highly
dispersed and were in a well-dened contact with ZnO. The latter is essential for the transfer of the species between ZnO and Cu. Moreover, it
is inferred in the literature [23] that a proximate contact between ZnO
and Cu enhances CO2, CO and H2 adsorption capacities due to the formation of additional sites at the interface. In our study, a direct monotonic
correlation between Cu content and methanol activity ratio was not observed when all catalysts were taken under consideration as a whole
set. Other factors besides Cu content need to be considered to account
for the activity trends of catalysts.
Therefore, for the observed behavior, the interaction between Cu
and ZnO plays a crucial role. The experiment on the electronic interaction between ZnO and the deposited Cu was not conducted directly, as
the observed activity (or even the methanol formation mechanism itself) is by large not on account of the (electronic) interaction

Fig. 6. Inuence of catalyst Cu content on its relative activity for methanol synthesis
(Eq. (1)).

44

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

mentioned. The vital interaction, nonetheless, relates to the (physical)


distribution and surface availability of both Cu and ZnO sites. In a typical
formate pathway, responsible for the CO2 hydrogenation to methanol,
the ZnO sites are crucial for CO2 adsorption, while the Cu sites for H2 adsorption, dissociation and the spillover to ZnO sites. Table 2 implies that
Cu surface area increased upon its deposition, yet, this occurred on account of ZnO sites. Thus, the interaction was the most favorable in the
case of 6.1 wt.% Cu, when a balance was obtained between the rates/
sites of/for CO2 and H2 adsorption and the mechanisms described.
Lower than this (Fig. 6), ZnO sites dominated, and the surface, available to H2 bonding was suboptimal, whereas on the other hand, going
past 6.1 wt.% Cu (Fig. 6) or the respective particle size of about 47 nm
(Fig. 7), caused Cu to commence blocking catalyst surface, leaving little
room for CO2 to adsorb, in turn, also preventing an effective H(s) spillover. Hence, it is this interaction that inevitably determined the optimal
catalyst material formulation of the so-designated Cu catalyst B.
3.2.2. Inuence of Cu crystallite size
Fig. 7 shows the dependence of the relative catalytic activity, calculated in the beginning of the methanol synthesis reactions, on the Cu
crystallite size after the reduction. Catalysts' activity is evidently decreasing with Cu particle size. At this point it has to be noted that the
catalytic activity of the prepared Cu/ZnO catalysts is indeed excellent,
as for example the relative catalytic activity of approximately 20
would imply a comparatively twentyfold greater conversion with regard to the commercial Cu/ZnO/Al2O3-based catalyst.
The highest relative catalyst activity toward the methanol synthesis
of 17.5 was obtained using sample B, since the crystallite size of the deposited Cu NP was low (46.6 nm) in comparison with the others. An increased size of the deposited Cu NP of 55.1 nm in sample C causes a
decrease of the relative methanol synthesis activity to the value of
10.9. The largest deposited Cu NP with the sizes of 62.4 nm was obtained
for sample E, causing a low relative methanol synthesis activity of 8.8 in
comparison to other prepared catalyst materials. In the work of
Natesakhawat et al. [6], the authors suggested that Cu particle size represents the predominant contributing factor for controlling the activity
of the catalysts for methanol synthesis reactions. Generally, smaller Cu
crystallites have a larger number of open planes and defect/edge sites,
which can bind strongly with intermediate species. Additionally, smaller Cu particles can lead to a larger interfacial area with ZnO support,
which also has an important role in overall catalyst activity [6]. Despite
the fact that the Cu crystallite size of the commercial catalyst, prepared
via co-precipitation method [1] is lower (17 nm) in comparison to those

of the prepared catalysts, the selectivity for methanol of the commercial


catalysts is signicantly lower. This nding points out that the dependence of the catalytic activity on the size of Cu crystallites is not a general characteristic, but a descriptor of a particular group of catalysts
only, obtained using a similar preparation method. In our study, however, a direct correlation between Cu particle size and methanol formation
is not observed when all the studied catalysts are taken under consideration. The latter strongly suggests that properties other than the average
Cu particle size affect the rate of methanol formation.
3.2.3. Inuence of composite preparation method
An improved development of the prepared catalyst surface seems to
be another key reason for the superior catalytic activity of the materials,
prepared using the photochemical method in comparison to the commercial catalysts. The oxides in the catalysts, prepared with the
commercial method, are inherently mixed, whereby Cu crystallites represent the main component (Cu/Zn N 1 (mol/mol)) and ZnO acts more
like a spacer rather than support. Frequently, Cu crystallites are in contact with several ZnO particles and the distance between them is
lowered. Thus, the exposed Cu surfaces are partially or completely covered with ZnO, whereas the neighboring Cu atoms tend to be in direct
contact, and hence, less accessible to the reacting molecules [24]. In
the catalysts, prepared using the photochemical method, the loading
of Cu is lower (Cu/Zn 1 (mol/mol)), while ZnO acts like a support, controlling the dispersion of Cu crystallites. The latter can easier bind with
the reacting molecules, leading to a better catalytic activity in comparison to the commercial analogues.
3.2.4. Inuence of Cu dispersion extent
The Cu crystallites deposited on ZnO support are dispersed on the
surface of ZnO support and consequently a large reactive surface area,
accessible for reaction molecules is provided. Small Cu crystallites are
commonly highly dispersed over the surface of ZnO particles and thus
cause a high overall surface area of the composites as well [18]. A high
dispersion of the Cu NP on ZnO is essential for achieving the desired catalytic activity, since it inhibits the sintering during methanol synthesis
and in the long run facilitates catalyst stability in addition to the initial
activity [25].
The dispersion of the Cu NP on the surface of ZnO nano-sheets is
strongly affected by the overall Cu content in the prepared catalysts.
In general, the catalyst B with the Cu content of 6.12 wt.%, which is comparably low, has Cu crystallites uniformly dispersed on the surface of
ZnO nano-sheets (Fig. 1c). On the other hand, the augmentation of the
Cu NP on the surface of ZnO nano-sheets (from 7.0 wt.%) deteriorates
the dispersion of Cu NP by means of forming agglomerated larger particles (Fig. 1f), and simultaneously, in turn decreases the relative methanol synthesis activity. The mechanism, responsible for this decrease of
relative activity, may as well be sought in a too extensive coverage of
the porous ZnO surface with Cu crystallites (Fig. 8a). Besides the
supporting function controlling the dispersion [26], ZnO acts as a reservoir of H2, providing additional hydrogen atoms via spillover effect to facilitate the hydrogenation of the adsorbed species on Cu NP (carbonate
and formate as the most prominent intermediate species) to form
methanol [27,28]. Thus, the extensive coverage of the ZnO support
with Cu inhibits the inherent interface contact between ZnO and Cu,
and consequently, also the spillover effect, leading to a decrease in catalytic activity [29]. Less Cu crystallites deposited on the surface of ZnO
(Fig. 8b) is therefore favorable for H2 spillover mechanism and may be
responsible for the increase in methanol production.
3.3. Catalytic selectivity, stability and deactivation of Cu/ZnO composites

Fig. 7. Cu crystallite size inuence on relative catalyst activity for methanol synthesis
(Eq. (1)).

The main products of the CO2 hydrogenation over the Cu/ZnO-based


catalysts, prepared with photochemical method were methanol, CO,
and H2O. Low amounts of methane, formaldehyde and dimethyl ether
were also detected as the products of reactions.

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

45

Fig. 8. STEM micrographs of catalyst E (a) and catalyst B (b) after reduction.

The catalytic activity of the prepared catalysts toward methanol production was determined at the overall pressure of 1 bar and at 523 K.
The reaction progress for each catalyst was monitored over the
timespan of 10 h, whereby the activity for methanol synthesis was evaluated at the beginning, middle and end of the process, while the proneness for methanol synthesis was expressed as a relative activity ratio,
using the activity of the commercial methanol synthesis catalyst for
benchmarking as follows:

Relative activity ratio Acatalyst =mcatalyst =Acommercial =mcommercial

Acatalyst and Acommercial are the average areas of peaks, describing


methanol species response (m/z = 32.04), for the catalyst, prepared
by photochemical method and commercial Cu/ZnO/Al2O3 catalyst, respectively. The relative activity ratio for methanol production as a function of reaction time is illustrated in Fig. 9a.

The catalytic activity of the synthesized Cu/ZnO dispersions is evidently superior in comparison to the conventional Cu/ZnO/Al2O3 catalysts during the whole reaction time. The relative activity ratio is
increasing over the time for all synthesized Cu/ZnO catalysts. This fact
can be explained by the strong observed decrease in the formation of
methanol upon using Cu/ZnO/Al2O3 catalyst over time, suggesting a
relatively rapid deactivation of the latter. Long-term methanol synthesis
experiments were performed to observe the progress of the catalysts deactivation process in CO2-rich syngas. The results show that Cu/ZnO/
Al2O3 catalyst showed a poor stability during methanol synthesis reactions, as the activity of such catalysts was reduced monotonously [30].
3.3.1. Catalytic stability and deactivation of Cu/ZnO composites by H2O
As reported in the literature, the deactivation of catalysts primarily
results from the formation of H2O, which is produced as a by-product
of CO2 hydrogenation (Eq. (2)) and the reverse watergas shift

Fig. 9. Relative methanol (a), water (b), carbon monoxide (c) and methane (d) production activity ratio as a function of reaction time.

46

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

(RWGS) reaction (Eq. (3)) as a subset of the overall methanol synthesis


process [31].
CO2 3H2 CH3 OH H2 O

CO2 H2 CO H2 O

According to the results, presented in Fig. 9b, the activity of the synthesized catalysts for H20 formation is lower in comparison with that of
the commercial catalyst, indicating that signicantly less H2O is produced as a by-product when the catalysts, prepared by the described
photochemical method, were applied. The latter can be seen as the underlying mechanism for a higher stability of the catalysts, prepared by
the employed photochemical method, since H20, which accumulates
in the pores in the available catalytic surface signicantly suppresses
the production of methanol [32]. Since the afnity of the Al2O3 carrier
for H2O is much stronger than several other substrates [20], a more extensive catalyst surface wetting occurs, and consequently, the blocking
of the CO2 hydrogenation surface sites of the commercial catalysts
should account for a lower catalytic activity during the methanol synthesis reactions in comparison with the prepared catalysts in this
work. In addition to the plugging of pores, Arena et al. [14] mention
that the presence of CuO, formed through the oxidation of Cu0,
may without a doubt deactivate the applied catalysts. The oxidized Cu
(CuO) can be in turn reduced back to Cu by H2 or CO
(CuO + H2 Cu + H20 and CuO + CO Cu + CO2). The work of Li
et al. [33] implies that specic structural changes of analogue catalysts
can accelerate the reduction of CuO back to its active state. The absence
of CuO in the catalysts B, C and E (Fig. 3) suggests that the deposition
preparation method may have promoted the rate of the reduction of
CuO by H2 or CO in the synthesized catalysts.
3.3.2. Catalytic stability and deactivation of Cu/ZnO composites by CO
In the work of Saito et al. [8] the authors reported that besides H2O,
CO suppresses the methanol synthesis over the Cu/ZnO-based catalysts,
while the inhibitory effect of H2O was reported to be much stronger
than that of CO. The desired synthesis reaction can be preceded by the
reverse watergas shift (RWGS) reaction (Eq. (3)) or it can alternatively
be formed in methanol decomposition reaction (Eq. (4)) [34].
CH3 OHCO 2H2

The calculated activity for CO formation was lower in comparison to


that of the commercial ones during the entire process time, meaning
that less CO was formed as a by-product when the synthesized catalysts
had been applied (Fig. 9c).
The methanol synthesis selectivity for the prepared catalysts was excellent during the whole process time (Fig. 9a), indicating that low
amounts of CO, which were formed as a by-product of the RWGS reaction (Eq. (3)), had no notable effect on methanol synthesis activity. In
contrast, a higher amount of CO more profoundly inuenced the activity
of the methanol formation for the commercial catalysts, since the latter
decreased with time and was also signicantly lower in comparison to
the activity of the prepared morphologically-structured analogs
(Fig. 9c). In the work of Sun et al. [35], the authors suggested, that the
presence of CO directly led to a gradual deactivation of their catalysts.
CO is known as a strong reducing agent, which can over-reduce the surface of ZnO particles and simultaneously achieve Cu to excessively undergo from the oxide state to neat metal, causing a non-desired
sintering of Cu crystallites. However, the deactivation, caused by CO, is
not as distinct as the activity decrease, instigated by H2O. Conversely,
the presence of products such as H2O, CO2 and methanol prevents the
surface over-reduction of catalysts, and consequently, also the migration of Cu clusters.

3.4. Mechanism of methanol formation


The activities of the commercial and synthesized catalysts for methanol, H2O and CH4 formation assist in getting a more detailed overview
of the mechanism of methanol formation. The latter was extensively investigated for several decades in order to nd out whether CO or CO2 is
the main reactant for methanol formation [36]. CO2 can be directly hydrogenated to methanol according to Eq. (2), while alternatively, CO2
is reduced to CO, according to Eq. (3), and the latter is subsequently hydrogenated to methanol (Eq. (5)).
CO 2H2 CH3 OH

It was already mentioned that H2O has an inhibiting effect on the


overall methanol formation, since it is competitively absorbed on the active sites of a catalyst and blocks CO2 hydrogenation sites. The activity of
the commercial catalysts for H2O formation was comparatively high
(Fig. 9b). The high concentrations of H2O led to a lower coverage of
the active sites with CO2, and in parallel, a higher coverage of the mentioned sites with CO, reaction intermediates being formed upon the reaction between CO and H2 [37,38]. Signicantly less H2O was produced
applying the prepared catalysts, suggesting that the hydrogenation of
CO2 must have been the dominant pathway for methanol production.
Additionally, the absorbed CO2 was also the source for the formation
of methane. It was proposed in the work of Klier et al. [39] that CH4 is
formed by the direct hydrogenation of CO2, not involving the RWGS
(Eq. (3)). The activities of the prepared Cu/ZnO catalysts for methane
formation were high during the whole reaction time (Fig. 9d), but decreased, which is another strong evidence of the high coverage of the actives sites with CO2 in prepared catalysts, being further hydrogenated
into either methane or (desired) methanol.
Periodic operation is generally not a suitable indicator for activity
tests, because it may show the amount of the active intermediate species for methanol formation, not the catalytic activity itself, in the present case, for example, the amount of the activated hydrogenated
species. Nonetheless, in the present study, the products of the reactions
at hand were analyzed by MS ex situ, not allowing the short-lived intermediates (e.g. formate species) to reach MS detector, in turn only determining stable reaction products, such as methanol, CO, and H2O.
Moreover, pulsing tests were utilized in order to simulate a realistic differing feedstock composition operation inherent to alternative carbon
capture and utilization (CCU) emerging processes (e.g. using exhaust
power plant ue gases, blast furnace gases, and biogas). Nonetheless,
to validate the ndings, presented in previous subchapters, a steadystate methanol synthesis operation was attempted as well, using the
same apparatus and conditions as for the pulsating measurements,
feeding CO2 and H2 in a stoichiometric ratio (Eq. (2)). The results revealed the steady state methanol formation activities of 0.93, 1.53,
1
for the catalysts A, B, C,
0.89, 0.74, 0.77, and 0.08 molmethanol kg1
cat. h
D, E, and the commercial Cu/ZnO/Al2O3 counterpart, respectively,
mirroring the obtained pulsed-measurement activities of Figs. 8 and 9.
The low absolute values of activities, nonetheless, have to be taken
into account to performing the experiments at low pressures (congure
1 bar as opposed to the conventional process pressure of approximately
50 bar). Nevertheless, even low-pressure data is usually taken as representative with regard to benchmarking activity and selectivity at high
pressures. Therefore, in relation to selectivity, similar steady-state yields
were obtained for CO, CH4 and H2O as well, that is at least in terms of
catalyst efciency ranking and general comparison trends.
4. Conclusions
The preparation of the Cu/ZnO materials by photochemical method
facilitates obtaining the catalysts with a higher activity towards methanol formation in comparison with those prepared using a conventional
co-precipitation method. Differing from the latter methods, where Cu

M.B. Pori et al. / Fuel Processing Technology 146 (2016) 3947

and ZnO particles can frequently be quite agglomerated, Cu particles


were supported on the surface of the ZnO carrier in the herein prepared
catalysts. A superior performance, exhibited by these catalysts, is therefore suggested to arise from the improved catalysts surface morphology,
allowing the Cu crystallites to more readily bind with the reacting molecules, leading to a better apparent catalytic activity in comparison to
the commercial catalysts. The catalytic activity of the prepared Cu/ZnO
catalysts was excellent during the whole methanol synthesis reaction
timespan examined, as for example the catalytic activity at the beginning of the reactions was up to 18-fold greater when benchmarked
against the commercial analogs. Methanol formation is favored over
catalysts with smaller Cu particles (4346 nm) when compared to
those with greater ones (5562 nm). The increase in the methanol formation over smaller Cu particles is proposed to arise from a larger interface contact between Cu and ZnO.
Besides the initial activity, the catalysts were also tested with respect
to their stability during methanol synthesis reactions. The prepared
catalysts deactivated to a lesser extent than the commercial ones. Preparing the catalysts with the photochemical method improved the catalytic stability by suppressing water and carbon monoxide formation.
The latter are formed as a by-product during the methanol synthesis
from CO2-rich feed and may cause the deactivation of catalysts.
Acknowledgment
The authors gratefully acknowledge the nancial support of the
Ministry of Higher Education, Science and Technology of the Republic
of Slovenia, and the Slovenian Research Agency (programs P2-0393
and P2-0152).
References
[1] A. Karelovic, A. Bargibant, C. Fernndez, P. Ruiz, Effect of the structural and morphological properties of Cu/ZnO catalysts prepared by citrate method on their activity
toward methanol synthesis from CO2 and H2 under mild reaction conditions,
Catal. Today 197 (2012) 109118.
[2] D. Wang, J. Zhao, H. Song, L. Chou, Characterization and performance of Cu/ZnO/
Al2O3 catalysts prepared via decomposition of M(Cu,Zn)-ammonia complexes
under sub-atmospheric pressure for methanol synthesis from H2 and CO2, J. Nat.
Gas Chem. 20 (2011) 629634.
[3] H. Ahouari, A. Soualah, A. Le Valant, L. Pinard, P. Magnoux, Y. Pouilloux, Methanol
synthesis from CO2 hydrogenation over copper based catalysts, React. Kinet. Mech.
Catal. 110 (2013) 131145.
[4] G.A. Olah, A. Goeppert, G.K.S. Prakash, Chemical recycling of carbon dioxide to methanol and dimethyl ether : from greenhouse gas to renewable, environmentally carbon neutral fuels and synthetic hydrocarbons, J. Org. Chem. 74 (2009) 487498.
[5] M. Behrens, A. Furche, I. Kasatkin, A. Trunschke, W. Busser, M. Muhler, et al., The potential of microstructural optimization in metal/oxide catalysts: higher intrinsic activity of copper by partial embedding of copper nanoparticles, ChemCatChem 2
(2010) 816818.
[6] S. Natesakhawat, J.W. Lekse, J.P. Baltrus, P.R. Ohodnicki, B.H. Howard, X. Deng, et al.,
Active sites and structureactivity relationships of copper-based catalysts for carbon
dioxide hydrogenation to methanol, ACS Catal. 2 (2012) 16671676.
[7] E. Samei, M. Taghizadeh, M. Bahmani, Enhancement of stability and activity of Cu/
ZnO/Al2O3 catalysts by colloidal silica and metal oxides additives for methanol synthesis from a CO2-rich feed, Fuel Process. Technol. 96 (2012) 128133.
[8] M. Saito, K. Murata, Development of high performance Cu/ZnO-based catalysts for
methanol synthesis and the watergas shift reaction, Catal. Surv. Asia 8 (2004)
285294.
[9] X. Zhai, J. Shamoto, H. Xie, Y. Tan, Y. Han, N. Tsubaki, Study on the deactivation phenomena of Cu-based catalyst for methanol synthesis in slurry phase, Fuel 87 (2008)
430434.
[10] M.V. Twigg, M.S. Spencer, Deactivation of copper metal catalysts for methanol decomposition, methanol steam reforming and methanol synthesis, Top. Catal. 22
(2003) 191203.

47

[11] D. Jingfa, S. Qi, Z. Yulong, C. Songying, W. Dong, A novel process for preparationj of a
Cu/ZnO/Al2O3 ultrane catalysts for methanol synthesis from CO2 + H2: comparison
of various preparation methods, Appl. Catal. A 139 (1996) 7585.
[12] M. Bitenc, M. Marinek, Orel Z. Crnjak, Preparation and characterization of zinc hydroxide carbonate and porous zinc oxide particles, J. Eur. Ceram. Soc. 28 (2008)
29152921.
[13] K.T. Jung, A.T. Bell, Effects of zirconia phase on the synthesis of methanol over
zirconia-supported copper, Catal. Lett. 80 (2002) 6368.
[14] F. Arena, G. Mezzatesta, G. Zafarana, G. Truno, F. Frusteri, L. Spadaro, How oxide
carriers control the catalytic functionality of the CuZnO system in the hydrogenation of CO2 to methanol, Catal. Today 210 (2013) 3946.
[15] K. Fujimoto, Y. Yu, Spillover effect on the stabilization of CuZn catalyst for CO2 hydrogenation to methanol, Stud. Surf. Sci. Catal. 3936 (1993).
[16] J.A. Moulijn, A.E. Van Diepen, F. Kapteijn, Catalyst deactivation : is it predictable ?
What to do ? 212 (2001) 316.
[17] V. Prakash, R.K. Diwan, U.K. Niyogi, Synthesis and characterization of copper oxide
nanopowders and their nanouids, J. Appl. Chem. 3 (2014) 10251030.
[18] V. Agarwal, S. Patel, K.K. Pant, H2 production by steam reforming of methanol over
Cu/ZnO/Al2O3 catalysts: transient deactivation kinetics modeling, Appl. Catal. A
Gen. 279 (2005) 155164.
[19] X. Guo, D. Mao, G. Lu, S. Wang, G. Wu, Glycinenitrate combustion synthesis of CuO
ZnOZrO2 catalysts for methanol synthesis from CO2 hydrogenation, J. Catal. 271
(2010) 178185.
[20] F. Arena, K. Barbera, G. Italiano, G. Bonura, L. Spadaro, F. Frusteri, Synthesis, characterization and activity pattern of CuZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol, J. Catal. 249 (2007) 185194.
[21] J. Agrell, H. Birgersson, M. Boutonnet, I. Melin-Cabrera, R.M. Navarro, J.L.G. Fierro,
Production of hydrogen from methanol over Cu/ZnO catalysts promoted by ZrO2
and Al2O3, J. Catal. 219 (2003) 389403.
[22] S. Mehta, G.W. Simmons, K. Klier, R.G. Herman, Catalytic synthesis of methanol from
CO/H2, J. Catal. 57 (1979) 339360.
[23] S. Natesakhawat, P.R. Ohodnicki, B.H. Howard, J.W. Lekse, J.P. Baltrus, C. Matranga,
Adsorption and deactivation characteristics of Cu/ZnO-based catalysts for methanol
synthesis from carbon dioxide, Top. Catal. 56 (2013) 17521763.
[24] I. Kasatkin, P. Kurr, B. Kniep, A. Trunschke, R. Schlgl, Role of lattice strain and defects in copper particles on the activity of Cu/ZnO/Al2O3 catalysts for methanol synthesis, Angew. Chem. Int. Ed. 46 (2007) 73247327.
[25] X.-M. Liu, G.Q. Lu, Z.F. Yan, J. Beltramini, Recent advances in catalysts for methanol
synthesis via hydrogenation of CO and CO2, Ind. Eng. Chem. Res. 42 (2003)
65186530.
[26] T. Fujitani, J. Nakamura, The effect of ZnO in methanol synthesis catalysts on Cu dispersion and the specic activity, Catal. Lett. 56 (1998) 119124.
[27] M.S. Spencer, Role of ZnO in methanol synthesis on copper catalysts, Catal. Lett. 50
(1998) 3740.
[28] R. Burch, S.E. Golunski, M.S. Spencer, The role of copper and zinc oxide in methanol
synthesis catalysts, J. Chem. Soc. Faraday Trans. 86 (1990) 26832691.
[29] S. Natesakhawat, P.R. Ohodnicki, B.H. Howard, J.W. Lekse, J.P. Baltrus, C. Matranga,
Adsorption and deactivation characteristics of Cu/ZnO-based catalysts for methanol
synthesis from carbon dioxide, Top. Catal. 56 (2013) 17521763.
[30] M. Kurtz, H. Wilmer, T. Genger, O. Hinrichsen, M. Muhler, Deactivation of supported
copper catalysts for methanol synthesis, Catal. Lett. 86 (2003) 7780.
[31] J. Wu, M. Saito, M. Takeuchi, T. Watanabe, The stability of Cu/ZnO-based catalysts in
methanol synthesis from a CO2-rich feed and from a CO-rich feed, Appl Catal A 218
(2001) 235240.
[32] S. Lee, A. Sardesai, Liquid phase methanol and dimethyl ether synthesis from syngas,
Top. Catal. 32 (2005) 197207.
[33] C. Li, X. Yuan, K. Fujimoto, Development of highly stable catalyst for methanol synthesis from carbon dioxide, Appl. Catal. A Gen. 469 (2014) 306311.
[34] A. Karelovic, P. Ruiz, The role of copper particle size in low pressure methanol synthesis via CO2 hydrogenation over Cu/ZnO catalysts, Catal. Sci. Technol. 5 (2015)
869881.
[35] J.T. Sun, I.S. Metcalfe, M. Sahibzada, Deactivation of Cu/ZnO/Al2O3 methanol synthesis catalyst by sintering, Ind. Eng. Chem. Res. 38 (1999) 38683872.
[36] Y. Hartadi, D. Widmann, R.J. Behm, Methanol formation by CO2 hydrogenation on
Au/ZnO catalysts effect of total pressure and inuence of CO on the reaction characteristics, J. Catal. 333 (2015) 238250.
[37] G. Liu, D. Willcox, M. Garland, H.H. Kung, The rate of methanol production on a copperzinc oxide catalyst : the dependence on the feed composition, J. Catal. 90
(1984) 139146.
[38] G. Liu, M. Willcox, M. Garland, H.H. Kung, The role of CO2 in methanol synthesis on
CuZn oxide: an isotope labeling study, J. Catal. 96 (1985) 251260.
[39] K. Klier, V. Chatikavanij, R.G. Hernman, G.W. Simmons, Catalytic synthesis of methanol from CO/H2, J. Catal. 74 (1982) 343360.

Вам также может понравиться