Sie sind auf Seite 1von 8

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

Available online at www.sciencedirect.com

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

Photocatalytic hydrogen evolution from aqueous


solutions of glycerol under visible light irradiation
Tatyana P. Lyubina a,1, Dina V. Markovskaya a,b,1,
Ekaterina A. Kozlova a,b,*, Valentin N. Parmon a,b,1
a
b

Boreskov Institute of Catalysis, Pr. Lavrentieva, 5, Novosibirsk 630090, Russia


Novosibirsk State University, St. Pirogova, 2, Novosibirsk 630090, Russia

article info

abstract

Article history:

Multiphase photocatalysts Pt/Cd1xZnxS/ZnO/Zn(OH)2 and single-phase photocatalysts

Received 1 July 2013

Pt/Cd1xZnxS were prepared by a two-step technique. The photocatalysts were character-

Received in revised form

ized by a wide range of experimental techniques: X-ray diffraction (XRD), high-resolution

30 July 2013

transmission electron microscopy (HRTEM) combined with energy-dispersive X-ray (EDX)

Accepted 5 August 2013

spectroscopy, low-temperature N2 adsorption/desorption, and UV/VIS spectroscopy. The

Available online 20 September 2013

photocatalytic activity was tested in a batch reactor in the reaction of hydrogen evolution
from aqueous solutions of glycerol under visible light irradiation (l > 420 nm). The highest

Keywords:

achieved photocatalytic activity was 449 mmol H2 per gram of photocatalyst per hour; the

Photocatalytic hydrogen evolution

highest quantum efficiency was 9.6% (l > 420 nm). The activity of the multiphase catalysts

Semiconductor photocatalysts

was shown to exceed that of the single-phase catalysts by a factor of 2.1, likely because of

Cd1xZnxS

the heterojunctions between sulfides, oxides and hydroxides.

Glycerol

Copyright 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

1.

Introduction

Nowadays the development of novel technologies for the


photocatalytic hydrogen production from renewable resources including water and biomass attracts a lot of attention
[1e4]. Hydrogen is a clean, environmental-friendly and
energy-rich fuel. Thus photocatalytic cleavage of water into
hydrogen and oxygen seems to be an important process for
sustainable energy production [5e7]. Unfortunately, the photocatalytic efficiency of the desirable process is quite low due
to electronehole pair recombination and the reverse reaction
of H2 with O2 on the semiconductor surface. The solution of
this problem is the use of electron donors (sacrificial agents),
which can react with formed oxygen or a photoinduced hole
[8e10]. The use of organic wastes or industrial pollutants as

sacrificial agents looks advantageous from a practical point of


view [11,12]. Recently glycerol (C3H8O3) is considered as a
prospective electron donor for the photocatalytic hydrogen
production because it is a by-product of the vegetable oil
transesterification into biodiesel [4,11].
The production of hydrogen via the photocatalytic water
splitting is possible when the conduction band (CB) edge of a
semiconductor is more negative than the electrochemical
hydrogen production potential and when the valence band
(VB) edge is more positive than the oxygen production
potential. TiO2 (Eg 3.2 eV) is a widely used photocatalyst for
hydrogen production from water solution of glycerol under UV
light irradiation [4,6,12e15]. However, the solar spectrum
contains only about 3e5% of the UV range irradiation [16].
Due to these limitations, the development of photocatalysts

* Corresponding author. Pr. Lavrentieva, 5, Novosibirsk 630090, Russia. Tel./fax: 7 (383) 3331617.
E-mail address: kozlova@catalysis.ru (E.A. Kozlova).
1
Tel./fax: 7 (383) 3331617.
0360-3199/$ e see front matter Copyright 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2013.08.031

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

working under visible light is of great interest for the use of the
solar light energy [17]. A quite narrow-band-gap of about
2.4 eV has the well-known semiconductor CdS. Unfortunately,
CdS is prone to undergo photocorrosion due to reactions with
photogenerated holes [18]. Some attempts have been made to
improve the stability and activity of cadmium sulfide. For
example, mixing of ZnS and CdS to form a Cd1xZnxS solid
solution has been shown to be an effective way to improve the
activity of CdS [19e22]. The band gap of the solid solution can
be continuously adjusted by changing its composition.
Another approach to the improvement of the photocatalytic
activity of sulfide photocatalysts is coupling the sulfide and
oxide nanoparticles [23]. Such multiphase composites can
combine the visible light responsibility of sulfides and stability
of oxides. Besides, the charge injection from the CB of the
narrow-band-gap semiconductor (CdS or CdS/ZnS) to the CB
of TiO2 or ZnO can lead to the efficient charge separation and
decreased electronehole recombination [16,24,25]. Recently
photocatalytic semiconductor materials containing heterostructures are widely used for the hydrogen production from
aqueous solutions of both inorganic [22,26e28] and organic
[12,29] electron donors.
In this research, we synthesized multiphase Pt/Cd1xZnxS/
ZnO/Zn(OH)2 and single-phase Pt/Cd1xZnxS photocatalysts
for the photocatalytic production of hydrogen from aqueous
solutions of glycerol under visible light irradiation. The
photocatalysts were synthesized using a two-step self-templated synthesis with Cd(OH)2 and Zn(OH)2 intermediates,
which has been proposed earlier for the CdS synthesis
[30,31]. Because the molar volume of sulfides is smaller than
that of hydroxides, this technique allows one to obtain materials with high porosity [32]. The photocatalysts were
characterized by the XRD, low-temperature N2 adsorptionedesorption, UVevis spectroscopy and TEM highresolution microscopy. The kinetic dependences of the
hydrogen evolution on the initial concentration of glycerol
and on pH were obtained under irradiation with the wavelength longer than 420 nm. It was shown that the activity of
the multiphase catalysts exceeds that of the single-phase
catalysts by a factor of 2.1.

2.

Experimental

2.1.

Catalyst preparation

The synthesis of composite photocatalysts was carried out


according to the method proposed by Bao et al. [30] for CdS;
Nasalevich et al. for CdS and CdS/Cd(OH)2 [31]; Kozlova and
Lyubina for Cd1xZnxS [32]. The synthesis goes through two
stages. The first stage is the formation of a hydroxide mixture:
0.1CdCl2 0.9ZnCl2 2NaOH 4 0.1Cd(OH)2 0.9Zn(OH)2
2NaCl.
(1)
The ratio of Cd to Zn equal to 1:9 was chosen on the basis of
our previous experiments [33] and results of Zhang et al.
[34,35].
The second stage is the formation of solid solutions of
cadmium and zinc sulfides:

14173

0.1Cd(OH)2 0.9Zn(OH)2 nNa2S 4 Cd1xZnxS 2NaOH, (2)


In this reaction, if n  1, the solid solution Cd0.1Zn0.9S is
formed. In contrast, if n < 1, the reaction results in a mixture of
Cd1xZnxS, Zn(OH)2 and Cd(OH)2.
A typical procedure was as follows. First, 10 ml of 100 mM
CdCl2 and 90 ml 100 mM ZnCl2 was placed in a 1000 ml glass
flask and mixed for 10 min. Then, 400 ml of 100 mM NaOH was
added dropwise under continuous stirring. After that, a proper
amount of a 100 mM Na2S solution was added in 15 min to
obtain a yellowish precipitate. Then, the suspension was
centrifuged and the resulting precipitate was washed thoroughly with distilled water five times. The catalysts were dried
at 70  C for 6 h.
The amount of added sodium sulfide differed for three
synthesized specimens:
Specimen A. The amount of 100 mM Na2S was 100 ml.
Specimen B. The amount of 100 mM Na2S was 50 ml.
Specimen C. The amount of 100 mM Na2S was 25 ml.
Note that specimens B and C were obtained at the lack of
Na2S and therefore consist of the mixture of sulfides and hydroxides. At that, a part of the constituent hydroxides can
transform to oxides during the drying.
After that, platinum was deposited on the surface of the
photocatalysts by the soft chemical reduction (SCR) technique. This technique was described in detail elsewhere [36].
Briefly, sulfide-based photocatalysts were impregnated with
an H2PtCl6 solution. Then a NaBH4 solution was added for
reducing Pt4 to Pt0. The content of platinum was 1 wt% for all
the photocatalysts.

2.2.

Catalyst characterization

Crystallographic structure of the photocatalysts was studied


by X-ray diffraction (XRD). The XRD patterns were obtained on
an X-ray diffractometer Xtra (Thermo, Switzerland) with a
CuKa source. The 2F angle ranged from 20 to 65 with a step
scanning of 0.05 for 3 s. The mass ratio of the phases was
calculated by means of the TOPAS package (General Profile
and Structure Analysis Software for Powder Diffraction Data,
Bruker AXS GmBH, Karlsruhe, Germany).
The elemental analysis was carried out on an ARL
ADVANTX 3.6 kW X-ray fluorescence spectrometer equipped
with an X-ray tube with a rhodium anode (voltage of 50 kV,
current of 40 mA). The samples were placed into a cartridge
covered with six films of spectrolene. Data were processed
using the QuantAS software.
The specific surface areas SBET, pore volumes and adsorptionedesorption curves of the photocatalysts were obtained
from the low-temperature N2 adsorptionedesorption (ASAP
2400, USA).
The surface morphology was studied by high-resolution
transmission electron microscopy (HRTEM) on a JEM-2010
electron microscope (JEOL, Japan) at the accelerating voltage
of 200 kV and a lattice resolution of 0.14 nm.
The diffuse reflectance spectra were recorded on a Perkin
Elmer UV/VIS spectrometer Lambda 35 with an integrating
sphere RSA-PE-20 (Labsphere, USA) in the wavelength range
from 200 to 1100 nm. Magnesium oxide was used as a reference material.

14174
2.3.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

Catalytic activity measurements

Cd1-xZnxS

Studies on the photocatalytic evolution of hydrogen from an


aqueous solution of glycerol were carried out by the method
described earlier [37]. A water suspension with a catalyst,
glycerol and NaOH was placed into a sealed thermostated
reactor and illuminated with a high-pressure mercury lamp
DRSh-1000 (1000 W, Russia) through a cutoff filter Yellow
Glass 11 (l  420 nm).
Before the reaction, the suspension was treated in an ultrasonic bath for 10 min and bubbled with argon for 20 min to
remove oxygen. The glycerol concentration varied from 50 to
1000 mM, the concentration of NaOH varied from 0 to 300 mM,
the catalyst concentration was 0.77 g L1, temperature was
20  C. The concentration of hydrogen was measured with a
gas chromatograph LChM-8 (Russia) equipped with a thermal
conductivity detector, a zeolite column and argon as the carrier gas.
To calculate the quantum efficiency of the photocatalysts,
the number of photons used for the hydrogen formation was
divided by the whole number of photons reached the sample
[38]. The participation of at least two photons is necessary to
generate one turnover of the hydrogen formation. The photon
flux, which was estimated by a technique proposed earlier
[32], was 7.8 mE min1. Therefore, the quantum efficiency F of
the sample can be found from the formula
F (%) (2  W/7.8)  100%,

(3)

where W is the rate of the hydrogen formation in mmol/min.


The quantum efficiencies calculated by formula (3) are
presented in Table 1.

3.

Results and discussion

3.1.

Catalyst characterization

3.1.1.

XRD, XRF and BET analysis

The XRD patterns of the photocatalysts are represented in


Fig. 1. The XRD pattern of specimen A exhibits only three
broad peaks at 28.8, 47.9 and 56.5 . These peaks correspond to
Cd1xZnxS solid solutions with x w 0.9 and the crystallite size
less than 2 nm. The XRD patterns of the specimens B and C are

Zn(OH)2

A
45

50

ZnO
55

60

C
B
A
20

40
2 , degrees

60

Fig. 1 e XRD patterns of photocatalysts A, B and C. Inset is a


detailed XRD pattern of specimen A.

more complicated: besides broad peaks of Cd1xZnxS (x is


close to 0.6e0.8), there are several sharp peaks attributed to
3-Zn(OH)2 with the crystallite size longer than 100 nm and
sharp peaks at 34.5, 36.3, 47.6, 56.7 and 62.9 of hexagonal ZnO
with the crystallite size about 35 nm. The peaks of Cd1xZnxS
are very broad and therefore the crystal structure (hexagonal
or cubic) cannot be determined accurately.
Note that specimens B and C contain pure hydroxide and
oxide of zinc, without cadmium traces. The solubility constant
of CdS is lower than that of ZnS (1.6  1028 and 2.5  1022,
respectively) and thus cadmium sulfide is formed from the
mixture of Cd(OH)2 and Zn(OH)2 first, and a part of zinc hydroxide remains unreacted when the lack of sodium sulfide is
used in stage 2. Drying at 70  C leads to the partial formation of
zinc oxide.
As described above, the ratio of Cd to Zn was equal to 1:9
and the quantity of Na2S was 100, 50, and 25% of the equimolar
quantity for specimens A, B, and C, respectively. The
elemental analysis confirmed that in the first case, Cd0.1Zn0.9S
was formed. Besides, the analysis has confirmed that the
molar percentage of sulfides was 50 and 25% for the photocatalysts B and C, respectively. Taking into account the mass
balance, the composition of photocatalyst B should be

Table 1 e Properties and photocatalytic activities of synthesized photocatalysts.


Catalyst
A
B

Phase composition

%, wt

CS,a nm

l, nm

SBET, m2/g

Vp, cm3/g

W0(H2),c mmol/min

F,b %

Cd0.1Zn0.9S
Cd0.2Zn0.8S
ZnO
3-Zn(OH)2
Cd0.4Zn0.6S
ZnO
3-Zn(OH)2

100
52
5
43
28
<1
w71

<2
<2
w35
>100
<2
w35
>100

430
453

228
50

0.24
0.18

0.131
0.273

3.3
7.0

442

44

0.21

0.230

5.9

a Crystallite size.
b Quantum efficiency.
c The rate was measured on platinized photocatalysts (1 wt% of Pt).

14175

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

Cd0.2Zn0.8S/ZnO/Zn(OH)2 while the composition of photocatalyst C is Cd0.4Zn0.6S/ZnO/Zn(OH)2. The values of x in


photocatalysts B and C agree well with the XRD data. The ratio
of ZnO and Zn(OH)2 was calculated using XRD data by means
of the TOPAS package. The results are represented in Table 1.
Table 1 shows that SBET of specimen A is higher than SBET of
specimens B and C. It may be caused by the difference in the
crystalline structure: the specimen A consists of only
Cd0.1Zn0.9S with a very small particle size, whereas the specimens B and C additionally contain zinc oxide and zinc hydroxide with the crystallite size of about 35 and 100 nm,
respectively.

3.1.2.

UVevis spectroscopy analysis

Diffuse reflectance spectra (Fig. 2a) were measured to determine the light absorption by photocatalysts A, B and C. The
spectrum of specimen A is typical of the single-phase
Cd0.1Zn0.9S solid solution. It is well known that Cd1xZnxS
solid solutions with x from 0.5 to 0.9 have the band gap energy
from 2.5 to 2.9 eV (430e495 nm) [20,21,32]; the absorption edge
of ZnO varies from 370 to 400 nm [39,40], whereas the Zn(OH)2
adsorption edge lies at about 250 nm [40]. The spectra of
composite photocatalysts B and C demonstrate a combination
of three spectra: Cd1xZnxS (range 450e600 nm), ZnO (range
350e450 nm) and Zn(OH)2 (range 350e250 nm). These spectra
confirm that specimen C consists mostly of Cd1xZnxS and
Zn(OH)2, whereas specimen B consists of Cd1xZnxS, ZnO and
Zn(OH)2. The relative intensities of the Cd1xZnxS, Zn(OH)2 and
ZnO spectra indicate that most of the surface is covered with a
sulfide solid solution.
We tried to estimate the band gap energy of Cd1xZnxS
nanoparticles in the composite photocatalysts. Because CdS
and ZnS are direct semiconductors, their absorption edges [41]
can be calculated using the Tauc function as shown in Fig. 2B.
The Tauc functions F(R)2(hv)2 versus hv are plotted in the range
of wavelengths from 420 to 600 nm because ZnO and Zn(OH)2
do not absorb light in this range. The band gap energy was
calculated by constructing a tangent to the curves and by
finding the X axis intercept for this tangent. The determined
adsorption edges are shown in Table 1. For specimen A, the
adsorption edge is 430 nm, which is in a good agreement with
the published data, which report that the adsorption edge of
Cd0.1Zn0.9S varies from 429 to 455 nm [21,32]. The evaluated
adsorption edges for specimens B and C are 452 and 442 nm,

respectively. This indicates that for these specimens, x is


lower than that for specimen A.

3.1.3.

3.2.

100

a
A
B
C

40
20

A
B
C

80
(F(R)h)

%R

60

Kinetic experiments

We conducted initial experiments on the hydrogen production from glycerol aqueous solutions under visible light irradiation with pure and platinized specimens A, B, and C. It was
shown that the rate of hydrogen evolution was equal to zero
on the non-platinized photocatalysts both in neutral and basic
media. It was also shown that no hydrogen was formed
with the use of platinized photocatalysts in neutral media.
Earlier, it was shown that the addition of NaOH increases
the adsorption of organic substrates on the CdS/ZnS

100
80

HRTEM and EDX analysis

The TEM images of specimens A and B are shown in Fig. 3aed.


In the case of specimen A, the analysis of periodic structures
allows identification of crystalline particles of a CdS/ZnS solid
solution with the characteristic size of a few nm. The sample
consists of hollow spheres with a diameter of approximately
30e40 nm and a wall thickness of 5e10 nm. However, the
coherently scattering domains do not exceed 2e5 nm, which
agrees well with the X-ray diffraction data. It was shown that
these structures are formed because the synthesis consisted
of two steps [32]. Fig. 3c and d shows that specimen B consists
of two types of particles: zinc oxide with the crystallite size
larger than 100 nm (zinc hydroxide in the conditions of the
TEM experiments transforms to the oxide) and Cd1xZnxS
crystalline particles with the characteristic size of a few nm.
The form of Cd1xZnxS nanoparticles is similar to that of the
specimen A: crystallites with coherently scattering domains
2e5 nm form hollow spheres (Fig. 3d).
The chemical composition of nanoparticles was additionally investigated by EDX technique. Fig. 3eeg represents that
in the case of specimen A and sulfide part of specimen B
lines of Cd, Zn and S are observed; in the case of oxide part of
specimen B lines of Zn and O are observed, the lines of S and
Cd are very low. Thus, EDX data confirms that the specimen B
consists of Cd1xZnxS and zinc oxide and hydroxide.
It should be noted that the proposed synthesis technique
provides a good contact between oxide and sulfide nanoparticles, which can likely increase the photocatalytic activity
of the synthesized materials.

60
40
20

0
250 300 350 400 450 500 550 600
, nm

0
2.6

2.8

3.0
E, eV

3.2

3.4

Fig. 2 e Diffuse reflectance spectra of photocatalysts A, B, C (a) and the Tauc plot for absorption edge determination (b).

14176

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

Fig. 3 e HRTEM images of specimens A (a, b) and B (c, d) with EDX (specimen A (e), specimen B sulfide part (f), specimen B
oxide part (g)).

25

amount of evolved H2, mol

photocatalysts [42]. The isoelectric points for ZnS and CdS are
in the range of 3.0e8.5, depending on crystalline structure [43],
so under basic conditions sulfide surface is negatively
charged. However, in basic media the substitution of surface
S2 by hydroxyl ions take place. Zinc hydroxide and cadmium
hydroxide have isoelectric points in the range 9.6e11.5 [44].
Thus, under basic conditions, the surface of the solid solution
(specimen A) and composites (specimens B and C) exists
mainly in ^ZneOH or ^ZneO and ^CdeOH or ^CdeO
forms [42]. We supposed that at pH w 13, glycerol, like glucose
[42], can form ions C3H7O that can compete with hydroxyls at
the Zn and Cd sites on the Cd1xZnxS surface, providing a
strong adsorption.
Fig. 4 shows that the rate of the hydrogen production under
basic conditions is quite slow at the beginning of the reaction
for all photocatalysts, but after about one hour, it grows and
becomes linear. The same results have been observed previously: the kinetic curves of the hydrogen production from
glycerol aqueous solutions had an apparent induction period
of 0.5e1.0 h [6,45,46]. According to the mechanism proposed
by Bowker et al. and de Oliveira Melo et al., on the first stage of

A
B

20

C
15
10
5
0
0

20

40

60

80

100

120

time, min
Fig. 4 e Kinetic curves of the photocatalytic hydrogen
evolution from aqueous solutions of glycerol in the
presence of synthesized photocatalysts with 1 wt% of Pt.
C(cat) [ 0.77 g LL1; T [ 20  C; C0 (glycerol) [ 100 mM, C0
(NaOH) [ 100 mM.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

the reaction of glycerol, CO evolves from C3H8O3 and poisons


the surface of the photocatalyst. Further, active oxygen species are formed after the reaction of water with photogenerated holes and oxidize CO to CO2. The removal of carbon
monoxide from the surface leads to an increase in the
hydrogen evolution rate [14,46].
Fig. 4 demonstrates that the photocatalytic activity of
multiphase specimens B and C is higher than the activity
of photocatalyst Pt/Cd0.1Zn0.9S. XRD, UVevis, and TEM
data confirm the presence of the heterostructures
Cd1xZnxS/ZnO/Zn(OH)2 in the specimens B and C. Zinc oxide
and solid solutions of zinc and cadmium sulfides form heterojunctions, which promote the charge separation. Earlier it has
been shown that the presence of heterojunctions in composites CdS/Cd1xZnxS [47] and TiO2/MoS2/graphene [29] is beneficial for the photocatalytic hydrogen evolution. Charge
separation mechanism in photocatalysts B and C is described
as follows: the potential of CB of sulfide is more negative than
that of ZnO, whereas the VB of ZnO is more positive than that
of Cd1xZnxS [48,49]. Under visible light irradiation, photogenerated electrons from the CB of Cd1xZnxS can transfer to
the CB of zinc oxide, and holes can transfer from the VB of ZnO
to that of the solid solution of zinc and cadmium sulfides. It is
well known that efficient charge separation decreases electronehole recombination and enhances the photocatalytic
activity [16,24,25,48]. Thus, multiphase photocatalysts
Pt/Cd1xZnxS/ZnO/Zn(OH)2 appear to be more active than the
single-phase photocatalyst Pt/Cd0.1Zn0.9S. The photocatalytic
activity of specimen C is lower than the activity of specimen B
likely because of the low content (about only 30%) of the visible
light responsible phase Cd0.4Zn0.6S.
The rate of the photocatalytic hydrogen evolution on the
best photocatalystdPt/Cd0.2Zn0.8S/ZnO/Zn(OH)2 (specimen
B)dwas 0.273 mmol min1 or 328 g1 h1. Earlier, the reported
rate on Pt/CdS was 107 mmol g1 h1 [46] under visible light
(l > 418 nm) and 388 mmol g1 h1 with the use of ZnO/ZnS
under the solar-simulated irradiation [48]. In our case, it is
better to compare our results with the results reported by de
Oliveira et al. [46] because the solar-simulated light source
used by Sang et al. [48] could produce the UV-irradiation as
well. Thus, the photocatalytic activity of synthesized multiphase photocatalysts is very high. The proposed synthesis
technique provides a good contact between the phases
because sulfides are formed directly from hydroxides. An

efficient contact between different semiconductor particles in


heterostructures allows one to achieve a good charge separation [48] leading to the high hydrogen evolution rate.
Fig. 5b shows the influence of the initial concentration of
glycerol on the initial rate of the hydrogen evolution. The
hydrogen evolution rate grows linearly with the increase in the
initial concentration of glycerol at low concentrations
(0e100 mM), then rises slowly until a maximum at 300 mM
(2.8 vol%) and then begins to fall at higher concentrations. The
further increase in the alcohol concentration leads to a decrease
in the activity, likely because of the saturation of the surface
active sites [45,46]. The highest photocatalytic activity achieved
at concentration 300 mM was equal to 449 mmol H2 per gram of
photocatalyst per hour, and the highest quantum efficiency
under visible light irradiation (l > 420 nm) was 9.6%. The rate of
the hydrogen production exceeds the values observed before
both under visible and solar-simulated light [46,48].
Earlier, the LangmuireHinshelwood mechanism was
observed for the photocatalytic hydrogen evolution from an
aqueous solution of organophosphorous dimethyl methylphosphonate [37]. It was shown that in the case of glycerol,
only the first part of the curve obeys this mechanism. We
approximated the dependence of the initial rate on the initial
concentration in the range of concentrations 0e300 mM by the
LangmuireHinshelwood equation:
W0 k 

W0, mol (H2)/min

W0, mol (H2)/min

0.18
0.12
0.06
0.00

0.4

0.24

50

100 150 200 250 300 350


C0(NaOH), m

K  C0
;
1 K  C0

(4)

where W0 is the initial H2 production rate (mmol H2/min), k is


the apparent rate constant (mmol/min), K is the adsorption
constant (mM1), and C0 is the initial concentration of glycerol
(mM). As seen from Fig. 5b, the experimental data for the small
glycerol concentration are well approximated by equation (4)
with the following parameters: k 0.53  0.05 mmol/min,
K 0.008  0.002 mM1.
Also, we studied the dependence of the initial rate of the
hydrogen evolution on the solution acidity. One can see that
without the addition of extra alkali, hydrogen does not evolve.
The rate of the hydrogen production significantly grows with the
addition of sodium hydroxide and achieves the maximum at the
NaOH concentrations from 50 to 100 mM, and then begins to fall.
As described above, the increase in pH leads to an enhancement
of organic substrate adsorption [42]. Simultaneously, with the
increase in the NaOH concentration, the redox potential of H/

0.36
0.30

14177

0.3
0.2

L-H

0.1
0.0

150 300 450 600 750 900 1050


C0(glycerol), mM

Fig. 5 e Initial rate of the hydrogen evolution in dependence on the initial concentrations of NaOH (a) and glycerol (b) (the
curve is the approximation by the LangmuireHinshelwood equation). Photocatalyst B with 1 wt% of Pt was used in all
experiments. C(cat) [ 0.77 g LL1; T [ 20  C; C0 (glycerol) [ 200 mM (a), C0 (NaOH) [ 100 mM (b).

14178

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

H2 becomes more negative, leading to the loss of the efficiency in


hydrogen evolution [50]. Thus, the dependence of the initial rate
of hydrogen evolution on the initial concentration of NaOH has a
maximum from 50 to 100 mM.

4.

Conclusions

A way for the synthesis of multiphase platinized photocatalysts Pt/Cd1xZnxS/ZnO/Zn(OH)2 has been proposed. It
was shown that these photocatalysts are more active in the
photocatalytic hydrogen evolution from aqueous solutions of
glycerol under visible light irradiation (l  420 nm) than most
common platinized single-phase Cd1xZnxS photocatalysts.
The highest achieved photocatalytic activity was 449 mmol H2
per gram of the photocatalyst per hour. The highest quantum
efficiency was 9.6% that exceeds the values reported before
[46]. High activity of multiphase systems may be caused by the
formation of heterojunctions between nanoparticles of
Cd1xZnxS, ZnO and Zn(OH)2, which result in the efficient
charge separation by minimizing the electronehole recombination. It is also found that the optimal pH for hydrogen
evolution is 12.5e13.5 while the optimal initial concentration
of glycerol is 2e4 vol%.

Acknowledgements
We gratefully acknowledge the support of Russian Federation
Department of Science and Education (Federal Target Program
Scientific and Educational Personnel, contract 8440; President Grant for the Leading Scientific Schools NSh-524.2012.3;
President Scholarship for Young Scientists SP-85.2012.1); RFBR
(grant #12-03-31104) and SB RAS (joint project #35). We also
acknowledge Drs. S.V. Cherepanova and E. Yu. Gerasimov for
help with obtaining and treating of XRD and TEM data.

references

[1] Zamaraev KI, Parmon VN. Potential methods and


perspectives of solar energy conversion via photocatalytic
processes. Catal Rev Sci Eng 1980;22:261e324.
[2] Graetzel M. Energy resources through the photochemistry
and catalysis. New York: Academic Press; 1983.
[3] Turner JA. Sustainable hydrogen production. Science
2004;305:972e4.
[4] Daskalaki VM, Kondarides DI. Efficient production of
hydrogen by photo-induced reforming of glycerol at ambient
conditions. Catal Today 2009;144:75e80.
[5] Jin ZL, Zhang XJ, Li YX, Li SB, Lu GX. 5.1% Apparent quantum
efficiency for stable hydrogen generation over eosinsensitized CuO/TiO2 photocatalyst under visible light
irradiation. Catal Commun 2007;8:1267e73.
[6] Gombac V, Sordelli L, Montini T, Delgado JJ, Adamski A,
Adami G, et al. CuO(x)-TiO2 photocatalysts for H2 production
from ethanol and glycerol solutions. J Phys Chem A
2010;114:3916e25.
[7] Park JH, Kim S, Bard AJ. Novel carbon-doped TiO2 nanotube
arrays with high aspect ratios for efficient solar water
splitting. Nano Lett 2006;6:24e8.

[8] Li YX, Wang JX, Peng ShQ, Lu GX, Li SB. Photocatalytic


hydrogen generation in the presence of glucose over ZnScoated ZnIn2S4 under visible light irradiation. Int J Hydrogen
Energy 2010;35:7116e26.
[9] Li YX, Lu GX, Li SB. Photocatalytic hydrogen generation and
decomposition of oxalic acid over platinized TiO2. Appl Catal
A 2001;214:179e85.
[10] Li YX, Me YZ, Peng SQ, Lu GX, Li SB. Photocatalytic hydrogen
generation in the presence of chloroacetic acids over Pt/TiO2.
Chemosphere 2006;63:1312e8.
[11] Yu J, Hai Y, Jaroniec M. Photocatalytic hydrogen production
over CuO-modified titania. J Colloid Interface Sci
2011;357:223e8.
[12] Yu J, Ran J. Facile preparation and enhanced photocatalytic
H2-production activity of Cu(OH)2 cluster modified TiO2.
Energy Environ Sci 2011;4:1364e71.
[13] Kondarides DI, Daskalaki VM, Patsoura A, Verykios XE.
Hydrogen production by photo-induced reforming of
biomass components and derivatives at ambient conditions.
Catal Lett 2008;122:26e32.
[14] Bowker M, Davies PR, Saeed Al-Mazroai L. Photocatalytic
reforming of glycerol over gold and palladium as an
alternative fuel source. Catal Lett 2009;128:253e5.
[15] Bahruji H, Bowker M, Davies PR, Saeed Al-Mazroai L,
Dickinson A, Greaves J, et al. Sustainable H2 gas production
by photocatalysis. J Photochem Photobiol A Chem
2010;216:115e8.
[16] Wu L, Yu JC, Fu X. Characterization and photocatalytic
mechanism of nanosized CdS coupled TiO2 nanocrystals
under visible light irradiation. J Mol Catal A 2006;244:25e32.
[17] Tristao JC, Magalhaes F, Corio P, Sansiviero MTC. Electronic
characterization and photocatalytic properties of CdS/TiO2
semiconductor composite. J Photochem Photobiol A Chem
2006;181:152e7.
[18] Chen J, Lin Sh, Yan G, Yang L, Chen X. Preparation and its
photocatalysis of Cd1xZnxS nano-sized solid solution with
PAMAM as a template. Catal Commun 2008;9:65e9.
[19] Youn HC, Baral S, Fendler JH. Dihexadecyl phosphate,
vesicle-stabilized and in situ generated mixed cadmium
sulfide and zinc sulfide semiconductor particles: preparation
and utilization for photosensitized charge separation and
hydrogen generation. J Phys Chem 1988;92:6320e7.
[20] Xing CJ, Yan W, Zhang YJ, Guo LJ. Band structure-controlled
solid solution of Cd1xZnxS photocatalyst for hydrogen
production by water splitting. Int J Hydrogen Energy
2006;31:2018e24.
[21] Zhang K, Jing D, Xing Ch, Guo L. Significantly improved
photocatalytic hydrogen production activity over Cd1xZnxS
photocatalysts prepared by a novel thermal sulfuration
method. Int J Hydrogen Energy 2007;32:4685e91.
[22] Li Q, Meng H, Zhou P, Zheng Y, Wang J, Yu J, et al. Zn1exCdxS
solid solutions with controlled bandgap and enhanced
visible-light photocatalytic H2-production activity. ACS Catal
2013;3:882e9.
[23] Kozlova EA, Kozhevnikova NS, Cherepanova SV, Lyubina TP,
Gerasimov EY, Kaichev VV, et al. Photocatalytic oxidation of
ethanol vapors under visible light on CdSeTiO2 nanocatalyst.
J Photochem Photobiol A 2012;250:103e9.
[24] Savinov EN, Gruzdkov YuA, Parmon VN. Suspensions of
semiconductors with microheterojunctions e a new type of
highly efficient photocatalyst for dihydrogen production
from solution of hydrogen sulfide and sulfide ions. Int J
Hydrogen Energy 1989;14:1e9.
[25] Gruzdkov YuA, Savinov EN, Korolkov VV, Parmon VN. Mixed
suspensions of semiconductors forming
microheterojunctions. A new type of highly efficient
photocatalysts: photocatalytic production of dihydrogen
from a water solution of sulfide ions in the presence of

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 8 ( 2 0 1 3 ) 1 4 1 7 2 e1 4 1 7 9

[26]

[27]

[28]
[29]

[30]

[31]

[32]

[33]
[34]

[35]

[36]

[37]

[38]

plasence of platinized ZnyCd1y/CuxS suspension. React


Kinet Catal Lett 1988;36:395e400.
Zhang J, Yu J, Jaroniec M, Gong JR. Noble metal-free reduced
graphene oxide-ZnxCd1exS nanocomposite with enhanced
solar photocatalytic H2-production performance. Nano Lett
2012;12:4584e9.
Zhang J, Yu J, Zhang Y, Li Q, Gong JR. Visible light
photocatalytic H2-production activity of CuS/ZnS porous
nanosheets based on photoinduced interfacial charge
transfer. Nano Lett 2011;11:4774e9.
Xiang Q, Yu J. Graphene-based photocatalysts for hydrogen
generation. J Phys Chem Lett 2013;4:753e9.
Xiang Q, Yu J, Jaroniec M. Synergetic effect of MoS2 and
graphene as cocatalysts for enhanced photocatalytic H2
production activity of TiO2 nanoparticles. J Am Chem Soc
2012;134:6575e8.
Bao NZ, Shen LM, Takata T, Domen K. Self-templated
synthesis of nanoporous CdS nanostructures for highly
efficient photocatalytic hydrogen production under visible
light. Chem Mater 2008;20:110e7.
Nasalevich MA, Kozlova EA, Lyubina TP, Vorontsov AV.
Photocatalytic oxidation of ethanol and isopropanol vapors
on cadmium sulfide. J Catal 2012;287:138e48.
Lyubina TP, Kozlova EA. New photocatalysts based on
cadmium and zinc sulfides for hydrogen evolution from
aqueous Na2S-Na2SO3 solutions under irradiation with
visible light. Kinet Catal 2012;53:188e96.
Lyubina TP, Kozlova EA. Unpublished results.
Zhang K, Jing D, Liu M, Guo L. Efficient photocatalytic H2
production under visible light irradiation over Ni doped
Cd1xZnxS microsphere photocatalysts. Catal Commun
2008;9:1720e4.
Zhang K, Jing D, Guo L. Effects of anions on the
photocatalytic H2 production performance of
hydrothermally synthesized Ni-doped Cd0.1Zn0.9S
photocatalysts. Int J Hydrogen Energy 2010;35:7051e7.
Kozlova EA, Vorontsov AV. Influence of mesoporous and
platinum-modified titanium dioxide preparation methods on
photocatalytic activity in liquid and gas phase. Appl Catal B
2007;77:35e45.
Kozlova EA, Vorontsov AV. Photocatalytic hydrogen
emission from aqueous solutions of organophosphorous
compounds. Int J Hydrogen Energy 2010;35:7337e43.
Braslavsky SE, Braun AM, ACassano AE, Emeline AV,
Litter MI, Palmisano L, et al. Glossary of terms used in

[39]

[40]

[41]

[42]

[43]
[44]

[45]

[46]

[47]

[48]

[49]
[50]

14179

photocatalysis and radiation catalysis (IUPAC


Recommendations 2011). Pure Appl Chem 2011;83:931e1014.
Rusdi R, Rahman AA, Mohamed NS, Kamarudin N,
Kamarulzaman N. Preparation and band gap energies of ZnO
nanotubes, nanorods and spherical nanostructures. Powder
Technol 2011;210:18e22.
Lin BC, Shen P, Chen SY. ZnO and 3-Zn(OH)2 composite
nanoparticles by pulsed laser ablation on Zn in water. J Phys
Chem C 2011;115:5003e10.
Xu X, Lu R, Zhao X, Xu S, Lei X, Zhang F, et al. Fabrication and
photocatalytic performance of a ZnxCd1xS solid solution
prepared by sulfuration of a single layered double hydroxide
precursor. Appl Catal B 2011;102:147e56.
Li Y, Gao D, Peng Sh, Lu G, Li Sh. Photocatalytic hydrogen
evolution over Pt/Cd0.5Zn0.5S from saltwater using glucose as
electron donor: an investigation of the influence of
electrolyte NaCl. Int J Hydrogen Energy 2011;36:4291e7.
Liu JC, Huang CP. Electrokinetic characteristics of some
metal sulfide-water interfaces. Langmuir 1992;8:1851e6.
Kosmulski M. Compilation of PZC and IEP of sparingly
soluble metal oxides and hydroxides from literature. Adv
Colloid Interface Sci 2009;152:14e25.
Lalitha K, Sadanandam G, Kumari VD, Subrahmanyam M,
Sreedhar D, Hebalkar NY. Highly stabilized and finely
dispersed Cu2O/TiO2: a promising visible sensitive
photocatalyst for continuous production of hydrogen from
glycerol:water mixtures. J Phys Chem C 2010;114:22181e9.
de Oliveira Meloa M, Almeida Silva L. Visible light-induced
hydrogen production from glycerol aqueous solution on
hybrid Pt-CdS-TiO2 photocatalysts. J Photochem Photobiol A
2011;226:36e41.
Yu J, Zhang J, Jaroniec M. Preparation and enhanced visiblelight photocatalytic H2-production activity of CdS quantum
dots-sensitized Zn1xCdxS solid solution. Green Chem
2010;12:1611e4.
Sang HX, Wang XT, Fan CC, Wang F. Enhanced
photocatalytic H2 production from glycerol solution over
ZnO/ZnS core/shell nanorods prepared by a low temperature
route. Int J Hydrogen Energy 2012;37:1348e55.
Kudo A, Miseki Yu. Heterogeneous photocatalyst materials
for water splitting. Chem Soc Rev 2009;38:253e78.
Peng Sh-Q, Peng Yu-J, Li Yu-X, Lu G-X, Li Sh-B. Photocatalytic
hydrogen generation using glucose as electron donor over Pt/
CdxZn1xS solid solutions. Res Chem Intermed
2009;35:739e49.

Das könnte Ihnen auch gefallen