Sei sulla pagina 1di 13

Nano Research

DOI 10.1007/s12274-016-1087-9

Photocatalytic reduction of CO2 with H2O over modified


TiO2 nanofibers: Understanding the reduction pathway
Anjana Sarkar1 (), Eduardo Gracia-Espino2, Thomas Wågberg2, Andrey Shchukarev1, Melinda Mohl3,
Anne-Riikka Rautio3, Olli Pitkänen3, Tiva Sharifi2, Krisztian Kordas3, and Jyri-Pekka Mikkola1,4 ()

1
Technical Chemistry, Department of Chemistry, Chemical-Biological Centre, Umeå University, SE-901 87 Umeå, Sweden
2
Department of Physics, Umeå University, SE-901 87 Umeå, Sweden
3
Microelectronics and Materials Physics Laboratories, Department of Electrical Engineering, University of Oulu, P.O. Box 4500, Oulu
FI-90014, Finland
4
Laboratory of Industrial Chemistry and Reaction Engineering, Johan Gadolin Process Chemistry Centre, Åbo Akademi University,
Biskopsgatan 8, Åbo-Turku FI-20500, Finland

Received: 25 June 2015 ABSTRACT


Revised: 27 March 2016 Nanosized metal (Pt or Pd)-decorated TiO2 nanofibers (NFs) were synthesized
Accepted: 1 April 2016 by a wet impregnation method. CdSe quantum dots (QDs) were then anchored
onto the metal-decorated TiO2 NFs. The photocatalytic performance of these
© Tsinghua University Press catalysts was tested for activation and reduction of CO2 under UV-B light. Gas
and Springer-Verlag Berlin chromatographic analysis indicated the formation of methanol, formic acid,
Heidelberg 2016 and methyl formate as the primary products. In the absence of CdSe QDs,
Pd-decorated TiO2 NFs were found to exhibit enhanced performance compared
KEYWORDS to Pt-decorated TiO2 NFs for methanol production. However, in the presence of
TiO2, CdSe, Pt-decorated TiO2 NFs exhibited higher selectivity for methanol, typically
photocatalysis, producing ~90 ppmg−1·h−1 methanol. The CO2 photoreduction mechanism is
CdSe quantum dots, proposed to take place via a hydrogenation pathway from first principles
CO2 photoreduction calculations, which complement the experimental observations.

1 Introduction reducing CO2 emissions. Increasing industrialization in


developing countries results in large-scale combustion
After realizing the after effects of the industrial of fossil fuels, such as coal, oil, and gas, contributing
revolution, increasing atmospheric CO2 concentration to further increases in CO2 levels in the atmosphere.
is now considered one of the main reasons for global Atmospheric levels of CO2 have increased by more
climate change. Along with methane (CH4), nitrous than 30% since 1700 B.C. [1]. This surge continues
oxide, halogenated compounds, and freons, CO2 to affect the Earth’s climate. Recently, researchers
constitutes a significant part of the greenhouse gases. predicted that the temperature of major cities across
Therefore, a lot of attention has been focused on the world will reach a climate point of no-return over

Address correspondence to Anjana Sarkar, sarkar.anjana@gmail.com; Jyri-Pekka Mikkola, jyri-pekka.mikkola@umu.se


2 Nano Res.

the next 20–30 years, or even sooner in tropical junctions with improved optical absorption and
countries [2]. The sooner we act, the better the climatic photocharge transfer/separation [35–38]. Anchoring
legacy inherited by future generations. Nature has noble metal nanoparticles on TiO2 photocatalysts also
established its own ways of recycling CO2 via the enhances their activity by trapping electrons on the
carbon cycle. It converts CO2 into various compounds metal surface, thus decreasing the recombination
that support the entire biological system on Earth. It rate of electron–hole pairs [6, 39–44]. Deposition of
is extremely important to maintain a balance of the semiconductor quantum dots (QDs) onto TiO2 for
amount of CO2 in the atmosphere in order to sustain photoinduced applications has also proved to be a
the eco-geo system. Thus, a permanent solution to the good strategy [6]. The conduction band (CB) of QDs,
CO2 problem must involve the transformation of CO2 such as CdSe, aligned on TiO2 becomes more negative
into other useful or non-toxic compounds. Upgrading due to electron confinement. This enables the injection
CO2 to reusable hydrocarbon resources would benefit of photo-excited electrons from the QDs into TiO2,
humans and the environment. which can subsequently be collected by an electrode
One promising route is artificial photosynthesis, or used to initiate a photocatalytic reaction. Recently,
which can be implemented via photoreduction of CO2 Wang et al. demonstrated CO2 photoreduction with
to produce hydrocarbons [3]. During the past few water using a CdSe QD-sensitized TiO2 heterostructured
decades, the capture of CO2 by photocatalytic processes catalyst using Pt as a co-catalyst under visible light
has attracted attention [4, 8–10]. Since CO2 is a rather [6]. They suggested that the combination of Pt and
inert and stable compound and its chemical reduction CdSe QDs is required for CO2 photoreduction into
is challenging, the capture and conversion of CO2 methane under their experimental setup.
requires high energy input as high temperature and In the present study, we report the photoreduction
pressure conditions are typically employed. However, of CO2 with water using TiO2 nanofiber (NF)-based
the photocatalytic process occurs under comparatively photocatalysts under UV light. The effect of metal
mild conditions with lower energy input, as the reaction nanoparticle (Pt or Pd) and CdSe QD co-catalysts on
is induced by either solar energy or other light the photocatalytic activity of TiO2 NFs is discussed.
sources. CO2 does not absorb in either the visible or Ab initio calculations are also performed, giving an
ultraviolet (UV) light regions ( = 200–900 nm), hence, insight into the photoreduction mechanisms, which
the process requires appropriate photosensitizers. support the experimental results.
Several studies have indicated the importance of
photocatalysts in enhancing the efficiency and 2 Results and discussion
selectivity of the photoreaction. Semiconductors such
as titanium dioxide (TiO2), zinc oxide (ZnO), zinc sulfide The TiO2 NFs were first loaded with 1 wt.% metal
(ZnS), hematite (-Fe2O3), tungsten oxide (WO3), nanoparticles (Pt and Pd) using a wet impregnation
strontium titanate (SrTiO3), cadmium selenide (CdSe), method. These catalysts are referred to as 1%Pt–TiO2
gallium oxide (Ga2O3), and zirconium dioxide (ZrO2) NFs and 1%Pd–TiO2 NFs, respectively. Finally, anchor-
are usually used as photocatalysts [4–11]. TiO2 has been ing of CdSe QDs on the TiO2-based catalysts (pristine
extensively studied with several reports stating that TiO2 NFs, 1%Pt–TiO2 NFs, and 1%Pd–TiO2 NFs) was
it is one of the most efficient photocatalysts when achieved. These catalysts are denoted as 1%CdSe–TiO2
considering cost, stability, and performance, and it NFs, 1%CdSe–1%Pt–TiO2 NFs, and 1%CdSe–1%Pd–
has been frequently applied in CO2 utilization [12–14]. TiO2 NFs.
Many studies have shown that CO2 can be reduced
2.1 Characterization of catalysts
with water vapor or other solvents by photocatalysts
[3–30]. Improvement of activity by nanoscale co-catalyst The amount of Pt, Pd, Cd, and Se in the catalysts was
materials has also attracted much attention [31–34]. determined by inductively coupled plasma-optical
Nanosized semiconductor co-catalysts with different emission spectrometry (ICP-OES). The results are listed
band structures in reference to TiO2 form hetero- in Table 1.

| www.editorialmanager.com/nare/default.asp
Nano Res. 3

The surface chemical composition of the catalysts to (CdOx) [47]. It could be also attributed to the
was analyzed by X-ray photoelectron spectroscopy decomposition of CdSe QDs in the presence of Pt
(XPS). The XPS results are summarized in Table 2. nanoparticles. Similarly, the Cd 3d spectrum for
The binding energies (BEs) for both Pt 4f7/2 and Pd 3d5/2 1%CdSe–1%Pd–TiO2 NFs shows an additional peak.
correspond to Pt(0) and Pd(0), respectively, showing However, in the 1%CdSe–1%Pd–TiO2 NF catalyst, the
slightly lower values than previously reported, probably Cd 3d5/2 peak at ~405 eV almost disappeared and a
due to the smaller size or elevated electron density on corresponding low binding energy peak at 403.6 eV
the metal nanoparticles [45]. In catalysts sensitized was more pronounced, as is clearly depicted in Fig. 1.
with CdSe QDs, the Cd 3d spectra showed typical A similar phenomenon is observed for the Cd 3d3/2
splitting of the Cd 3d core levels into 3d5/2 (404.7 ± peaks for both catalysts. These results indicate that
0.4 eV) and 3d3/2 (~411.5 eV) components by spin-orbital CdSe QDs are more stable in the presence of Pt than
coupling with a splitting magnitude of ~6.7 eV [46]. Pd nanoparticles.
Interestingly, the Cd 3d spectra are different in The size and morphology of the modified TiO2 NFs
the presence of Pt and Pd nanoparticles. For the were analyzed using transmission electron microscopy
1%CdSe–1%Pt–TiO2 NF catalyst, the Cd 3d5/2 core level (TEM), high-resolution TEM (HRTEM), scanning
shows the presence of an additional low intensity TEM (STEM), and field emission scanning electron
peak at 403.3 eV, which most probably corresponds microscopy (FESEM). TEM analysis revealed the size

Table 1 The actual amount of Pd, Pt, Cd, and Se in the samples from ICP-OES analysis
Photocatalyst Pd (mg·g–1) Pt (mg·g–1) Cd (mg·g–1) Se (mg·g–1)
1%Pd–TiO2 NFs 7.7
1%Pt–TiO2 NFs 6.6
1%CdSe–TiO2 NFs 1.4 <1
1%CdSe–1%Pd–TiO2 NFs 7.7 1.4 <1
1%CdSe–1%Pt–TiO2 NFs 6.6 1.4 <1

Table 2 Binding energies and atomic concentrations (at.%) of the TiO2-based catalysts from XPS analysis
1%Pd–TiO2 1%CdSe–TiO2 1%CdSe–1%Pt– 1%CdSe–1%Pd– Peak
TiO2 NFs 1%Pt–TiO2 NFs
NFs NFs TiO2 NFs TiO2 NFs assignment
458.5 eV; 458.8 eV; 458.9 eV; 458.6 eV; 458.5 eV; 458.6 eV; TiO2
Ti 2p3/2
24.98 at.% 20.17 at.% 24.21 at.% 21.35 at.% 18.17 at.% 19.22 at.%
334.5 eV; Pd (0)
335.0 eV; 0.33 at.%
0.63 at.% 335.1 eV;
0.18 at.%
Pd3d5/2 Not detected Not detected Not detected Not detected
336.1 eV; 335.9 eV; Pd (II)
0.25 at.% 0.05 at.%
337.5 eV; 337.1 eV; Pd (II)
0.22 at.% 0.03 at.%
70.7 eV; 0.17 70.5 eV; 0.15
Pt(0)
at.% at.%
Pt4f7/2 Not detected Not detected Not detected Not detected
71.9 eV; 0.05
Pt(II)
at.%
403.6 eV; 403.3 eV;
CdOx
0.02 at.% 0.09 at.%
Cd 3d5/2 Not detected Not detected Not detected
404.4 eV; 405.0 eV; 404.7 eV;
CdSe
0.06 at.% 0.14 at.% 0.03 at.%

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


4 Nano Res.

distribution of spherical Pt and Pd nanoparticles to


be highly monodispersed, centered at 2.0 ± 0.5 and
5.0 ± 1.1 nm, respectively (see Fig. S2 in the Electronic
Supplementary Material (ESM)). After anchoring of
the CdSe QDs (2.1–2.3 nm and full width half maximum
(FWHM) = 32 nm (max = 459 nm)), we did not observe
any significant change in the cluster size of the noble
metals. Unfortunately, due to their small size the CdSe
QDs did not exhibit proper contrast in the TEM
images, complicating their characterization. HRTEM
and STEM analyses of these samples gave a better
understanding of their morphology (Fig. 2). HRTEM Figure 1 XPS spectra of the Cd 3d core level for (a) 1%CdSe–
and STEM images clearly show that in the 1%CdSe– TiO2, (b) 1%CdSe–1%Pt–TiO2 NFs, and (c) 1%CdSe –1%Pd–
1%Pd–TiO2 NFs, particles were formed in two different TiO2 NFs.
size ranges; the smaller nanoparticles are less than
1.5 nm in diameter and most of them are ~0.7 nm,
while the larger are in the range of ~5 nm. Due to the
strong background of crystalline TiO2 in the HRTEM
images, which makes it difficult to recognize the
borders of the nanoparticles (especially the smaller
ones), it is more reliable to use the STEM images for
actual size measurements. For 1%CdSe–1%Pt–TiO2,
a more homogeneous particle size distribution was
observed. Here, particles are slightly bigger and those
with diameters of below 1 nm are rarely observed.
Although it is not possible to distinguish different types
of nanoparticles based solely on electron microscopy,
it seems plausible to suggest that CdSe QDs are
represented by the small nanoparticles in the range
of 0.7–1.5 nm, while the larger nanoparticles in the
1%CdSe–1%Pd–TiO2 NF samples are Pd. The fact that
some particles with the same size display differences
in contrast in the STEM image of the 1%CdSe–1%Pt– Figure 2 HRTEM images of (a) 1%CdSe–1%Pd–TiO2 NFs and
TiO2 NFs clearly supports co-existence of different (c) 1%CdSe–1%Pt–TiO2 NFs, and STEM images of (b) 1%CdSe–
types of nanoparticles, where a higher contrast is 1%Pd–TiO2 NFs and (d) 1% CdSe–1%Pt–TiO2 NFs.
expected for Pt particles due to their significantly higher
mass. Nevertheless, the small size of the particles and which correspond to the characteristic anatase phase
the charging effect in the TiO2 rods make it very of TiO2 NFs.
difficult to perform chemical mapping, which renders
2.2 Photocatalytic reduction of CO2
unambiguous assignment impossible. It is not there-
fore possible to identify particular nanoparticles. The photocatalytic reduction of CO2 with water into
Topographical images of the samples obtained by hydrocarbons using TiO2-based photocatalysts has
FESEM are displayed in Fig. S5 in the ESM. been widely studied [7], revealing that the reduction
The Raman spectra (see Fig. S2 in the ESM) of the mechanism is mainly initiated by the formation
samples showed four distinct peaks detected at 147 of two important radicals: •H (hydrogen radical)

(Eg), 399 (B1g), 517 (A1g + B1g), and 639 cm−1 (Eg), and •CO2 (carbon dioxide anion radical). Researchers

| www.editorialmanager.com/nare/default.asp
Nano Res. 5

have reported a diverse range of products formed upon reactions at lower energies [38, 50]. Due to the low
photoreduction of CO2, such as methane, methanol, energy of UV-B lamps, it is possible to monitor the
carbon monoxide, hydrogen, methyl formate, formic formation of initial products during the reaction. In
acid, ethylene, ethanol, and formaldehyde [48]. all cases, under our experimental conditions, liquid
The CO2 photoreduction performance of our TiO2- products condensed in the cold trap revealed the for-
based catalysts (i.e. pristine TiO2 NFs, 1%Pt–TiO2 NFs, mation of methanol, formic acid, and methyl formate
1%Pd–TiO2 NFs, 1%CdSe–TiO2 NFs, 1%CdSe–1%Pt– as the major products, as well as other hydrocarbons
TiO2 NFs, 1%CdSe–1%Pd–TiO2 NFs) was tested under in low yields. Figure 3 shows the products (normalized
UV-B light. It has been reported that metal clusters area%) produced for each catalyst system. Formic
decorated on TiO2 act as electron trapping centers, acid was the major product with 1%Pt–TiO2 NFs, and
leading to charge separation due to newly formed low yields of methanol and methyl formate were also
Schottky barriers at the metal/semiconductor interface obtained. On the other hand, 1%Pd–TiO2 NFs formed
[49]. Upon light irradiation, electrons and holes are methanol, formic acid, and methyl formate as the
generated. The electrons formed migrate to the metal major products. In the case of methanol formation,
surface and their back injection is inhibited by the our results agree with those previously reported in
Schottky barriers, thus limiting recombination. The the literature; namely, Pd-decorated TiO2 nanoparticles
electron rich metal surface acts as a reduction site, are preferred over Pt, as suggested by Ishitani et al. [51].
while oxidation takes place at the photogenerated For 1%CdSe–1%Pt–TiO2 NFs, the presence of CdSe
holes. Consequently, introducing Pt or Pd can influence QDs drastically enhances not only the formation of
the CO2 reduction reaction rate. In addition, anchoring methanol, but also improves overall product formation.
CdSe QDs affects the electron distribution on the TiO2 The rate of reaction increases, thus increasing methanol
surface [38]. The CB of bulk CdSe exhibits a slightly formation. Formation of acetic acid and ethanol in
negative potential compared to TiO2. However, due very low yields was also observed. Meanwhile, in the
to quantum confinement, the conduction band of CdSe case of 1%CdSe–1%Pd–TiO2 NFs, the formation of
QDs shifts to a more negative potential, enabling methanol decreased and yield of formic acid increased
electron injection into TiO2 [38]. The electron transfer approximately 3-fold. The decrease in photoactivity
rate is size-dependent and is enhanced as the size of towards methanol formation can be attributed to the
the QDs decreases. The time scale for electron transfer instability of CdSe QDs in the presence of Pd metal, as
from excited CdSe QDs to TiO2 has been reported to be mentioned earlier. Under our experimental conditions,
around 2–50 ps, which facilitates the photocatalytic the combination of Pt and CdSe QDs was found to be
activity of TiO2 for reduction reactions [6]. Additionally, the most efficient for methanol formation. Table 3
CdSe–TiO2 heterostructures are known to harvest summarizes the yields of methanol and methyl formate
visible light energy and facilitate photon-induced for each catalyst.

Figure 3 Products formed for each catalyst (A = methanol, B = methyl formate, C = formic acid, D = other low yield products).

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


6 Nano Res.

Table 3 Methanol and methyl formate yields from photoreduction that there is no general consensus on the reaction
of CO2 with water using TiO2-based catalysts mechanism, but most studies have indicated the
Methanol Methyl formate existence of two reaction pathways, depending upon
Photocatalyst concentration concentration
the rates of CO2 hydrogenation and deoxygenation
(ppm·g–1·h–1) (ppm·g–1·h–1)
[7]. If CO2 hydrogenation is prevalent, the formation
1%Pt–TiO2 NFs 1.61 ± 0.05 11.62 ± 0.5
of formic acid, formaldehyde, and finally methanol
1%CdSe–1%Pt–
90.22 ± 0.7 225.4 ± 1.0 will be observed. Alternately, carbon monoxide and
TiO2 NFs
methane are formed when deoxygenation is faster
1%Pd–TiO2 NFs 14.21 ± 0.04 14.79 ± 0.01
than hydrogen addition. On the basis of this infor-
1%CdSe–1%Pd–
4.63 ± 0.04 13.11 ± 0.3 mation, we propose that under our experimental
TiO2 NFs
conditions CO2 photoreduction with water over TiO2
1%TiO2 NFs 3.07 ± 0.05 53.42 ± 0.01
NFs mainly occurs via hydrogenation reactions, as
the intermediates and products are consistent with
Sensitizing TiO2 NFs with semiconductor CdSe QDs
those species observed experimentally (methanol,
not only enhances the yield, but can also influence the
formic acid, and methyl formate). The proposed reaction
selectivity of the reaction. It is important to mention
pathways are stated in Eqs. (1)–(12), which have been
that CdSe–TiO2 NFs did not show any influence on
previously observed in similar studies over a diverse
product yield compared to pristine TiO2 NFs. These
range of surfaces [52, 53].
series of experiments clearly suggest a synergy between
the metal (Pt or Pd) nanoparticles and CdSe QDs over TiO2 + he– (CB) + h+ (VB) (1)
TiO2 nanofibers and demonstrated their enhanced
H2O+ h  OH + H
+ +
(2)
photocatalytic activity towards CO2 reduction. Blank
experiments were also carried out to ensure that the H + e  H
+ –
(3)
products formed are solely due to photoreduction of CO2(g)  CO2 (4)
CO2. Under similar experimental conditions, blank
filter paper without catalyst was illuminated, and CO2 + H HCOO* (5)
experiments in the dark with catalyst were also carried HCOO + H HCOOH (formic acid) (6)
out. Furthermore, an additional blank test was
HCOOH + H H2COOH (7)
performed using nitrogen gas instead of CO2. Hydro-
carbons were not detected in any of these cases, and H2COOH  CH2O+OH (8)
only water was condensed in the cold trap. Under
CH2O + OH + H CH2O + H2O(g) (9)
our experimental conditions, no detectable amount of
gaseous products was observed. CH2O + H CH3O (10)
CH3O+ H  CH3OH(g) (methanol) (11)
2.3 Theoretical description of CO2 reduction reaction
on TiO2 HCOOH + CH3OH  H2O + HCOOCH3 (12)
(methyl formate)
In the previous sections, we described our experimental
observations for the CO2 photoreduction process over where  indicates adsorbed species, and hrepresents
TiO2 NFs, however, the influence of Pt and Pd is still a photon of appropriate wavelength. Here the first
unclear. We complemented our experimental observa- three reactions (Eqs. (1)–(3)) involve the absorption

tions with theoretical studies using DFT. Thus, we of a photon, creating an electron (e )–hole (h+) pair
performed diverse calculations to study the reaction in the CB and valence band (VB), respectively. This
mechanism and increase understanding of the observed electron–hole pair will further react with previously
results. As previously mentioned, CO2 reduction is a adsorbed water molecules, producing OH and H
complex reaction, and a vast number of mechanisms intermediates (Eqs. (2) and (3)). The following reactions
have been proposed [7]. It should be mentioned in Eqs. (4)–(12) are based on a fast hydrogenation

| www.editorialmanager.com/nare/default.asp
Nano Res. 7

model of the CO2 molecule over the TiO2 surface, the catalyst, as observed experimentally.
where the consecutive addition of 6 H atoms (H The calculated reaction energy for CO2 reduction is
intermediates) generates a diverse range of organic depicted in Fig. 4(b). As observed, CO2 adsorption at
compounds, similar to those observed experimentally. a Ti5f site on the anatase TiO2 (101) surface exhibits a
We have only calculated reaction energy (ΔER) during stable configuration with absorption energy (ΔEads)
CO2 reduction, as shown in Eqs. (4)–(11), since these equal to −0.17 eV, in good agreement with previous
reactions are specifically involved in the hydrogenation reports [54, 55]. The first hydrogenation step (Eq. (5))
of the CO2 molecule. The results are depicted in Fig. 4. results in a transition state of relatively high energy
Figure 4(a) illustrates the reaction pathway on the (ΔER5 = 2.1 eV) that most probably rapidly reacts to
anatase TiO2 (101) surface based on a fast hydro- form the first stable intermediate, formic acid (Eq. (6)).
genation mechanism. As indicated in Eq. (4), the As observed in Fig. 4(b), the hydrogenation of formic
reaction is initiated by the adsorption of a CO2 acid (Eq. (7)) exhibits the largest ΔER, equal to 2.4 eV,
molecule over a 5-fold coordinated Ti atom (Ti5f, see and is thus considered the rate-limiting step. It is
also Fig. S4.1(a) in the ESM). It is possible to observe important to note that, due to the large amount of
that the CO2 molecule is preferentially adsorbed on a energy required for formic acid hydrogenation to
tilted configuration, facing one O atom directly to the proceed, formic acid could actually desorb as a by-
Ti5f with a CO2–Ti5f distance equal to 2.6 Å. These product. However, if the larger adsorption energy of
results are in good agreement with those reported in formic acid is observed, the opportunity of continuing
previous studies [53]. The following steps involve CO2 reduction is increased substantially. These results
hydrogenation reactions, except for that shown in are in excellent agreement with our experimental
Eq. (8), where the reaction involves the dissociation of observations, where a considerable amount of formic
a H2COOH intermediate into formaldehyde (CH2O) acid is observed during CO2 reduction.
and hydroxyl (OH). Then, by further addition of H The driving force for CO2 photoreduction over TiO2
intermediates, the reaction proceeds until methanol is assumed to be the difference between the redox
is obtained (CH3OH). Finally, Eq. (12) shows the potential corresponding to the VB of TiO2 at the
condensation reaction of methanol and formic acid to surface under irradiation (~3.2 eV) [56]. According to
produce methyl formate. Since it does not involve a our calculations, illumination with light provides the
hydrogenation step, it is not included in Fig. 4(b). It is potential needed for the reaction to proceed and,
important to point out that here we consider the therefore, photoreduction of CO2 over anatase TiO2
photochemical reactions (Eqs. (1)–(3)) as the source of (101) is thermodynamically favored. Once the hydro-
adsorbed hydrogen (H), thus accumulating OH genation of formic acid has been achieved, the
intermediates at the TiO2 surface that could deactivate following chemical reactions (Eqs. (8)–(11)) proceed

Figure 4 (a) Geometrically optimized structures of the diverse range of intermediates along the CO2 reduction pathway performed over
the anatase TiO2 (101) surface. The molecules and intermediates inside the circles indicate additions to the system. (b) ΔER during CO2
reduction, as described in Eqs. (4)–(11). In this case, the limiting step is the hydrogenation of formic acid (Eq. (7)). * indicates adsorbed
species. Color code in (a): blue (H), yellow (C), red (O), and gray (Ti).

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


8 Nano Res.

smoothly due their smaller ΔER. Similarly, the relatively modified TiO2 NFs was demonstrated under a low
low ΔER for the reaction in Eq. (10) could explain energy UV-B light source. TiO2 NFs decorated with
why formaldehyde is not observed experimentally, Pt and Pd nanoparticles lead to the formation of similar
since its low hydrogenation energy (ΔER10 = 1.1 eV) products, but exhibited selectivity towards different
facilitates methanol production. Finally, the condensa- products. Decorating TiO2 NFs with Pd metal is
tion reaction of methanol and formic acid (Eq. (12)) preferred over Pt for methanol formation. Anchoring
exhibits ΔER12 = −0.14 eV, indicating that this reaction CdSe QDs on Pt-decorated TiO2 NFs results in a drastic
is also favored. enhancement in methanol formation and also enhances
Identification of the rate-limiting step permits the overall yield of products. Nevertheless, CdSe
deeper investigation into the effect of Pt and Pd nano- QDs anchored on Pd-decorated TiO2 NFs exhibited
particles on the CO2 photoreduction process. As stated selectivity towards formic acid and decreased the
previously, metals such as Pt or Pd act as electron yield of methanol due to decomposition of CdSe QDs
traps, modifying the net charge of TiO2. Considering in the presence of Pd nanoparticles. The synergistic
the current limitations and computational costs of effect of CdSe QDs and Pt was found to be optimal
actual ab initio approaches, we implicitly included the for methanol formation under our experimental con-
effect of Pt and Pd on the TiO2 surface by varying the ditions. The photoreduction of CO2 is proposed to take
total charge of the system. The net charge on anatase place via a hydrogenation pathway, forming various
TiO2 (101) was modified from −0.4 to + 0.4 e–, as products. Modified TiO2 displayed enhanced activity
suggested in previous publications [57]. We then towards CO2 photoreduction, which was demonstrated
calculated the change in adsorption energy of formic experimentally and using DFT calculations.
FA
acid (ΔEads ). Our results show that the adsorption
energy changes linearly with the net charge of the
4 Experimental
system (see Fig. S4.3 in the ESM), where a positively
FA
charged surface exhibits a stronger ΔEads by ~0.12 eV 4.1 Materials
at +0.4 e– when compared with neutral systems. Thus,
improving the interaction of formic acid may benefit Titanium dioxide (Sigma-Aldrich, <25 nm anatase,
hydrogenation. On the other hand, a negatively 99.7%) powder was utilized as received. Water purified
FA
charged system exhibits a weaker ΔEads (a decrease of through a Millipore system was used. Precursors Pt-
~0.12 eV at −0.4 e–), indicating faster desorption of and Pd-acetylacetonate were also received from Sigma
formic acid and inhibition of further reduction reactions. Aldrich, as well as CdSe QDs 2.1–2.3 nm in size.
Similar results were observed when the rutile TiO2
4.2 Catalyst synthesis
(110) surface was used as the catalyst, see Fig. S4.3 in
the ESM. These results agree qualitatively with our Synthesis of sodium titanate (Na2TiyO2y+1) NFs was
experimental observations, where the presence of achieved via hydrothermal treatment of anatase TiO2
metal nanoparticles generates a slightly positive TiO2 in 10 M NaOH aqueous solution for 24 h at 175 °C
surface, therefore increasing the adsorption energy of [49]. Furthermore, the resulting material was treated
formic acid and improving CO2 reduction. We also with 0.1 M hydrochloric acid to exchange sodium
FA
observed a larger ΔEads on anatase TiO2 (101) (−0.87 eV) ions with protons in the NFs. The product was
than on rutile TiO2 (101) (−0.63 eV), which suggests washed and dried at 70 °C, followed by calcination at
that anatase TiO2 (101) might act as a better catalyst for 600 °C in air for 12 h to form TiO2 NFs.
formic acid hydrogenation and subsequent production Loading of Pt and Pd nanoparticles on TiO2 nano-
of methanol and other related products. wires was achieved by means of a wet impregnation
method using Pt- and Pd-acetylacetonate precursors,
respectively. In a typical procedure, 1 wt.% precursor
3 Conclusions
salt was dissolved in 50 mL acetone (VWR) and mixed
In summary, photocatalytic reduction of CO2 using with 1.0 g TiO2 NFs. The suspension was sonicated

| www.editorialmanager.com/nare/default.asp
Nano Res. 9

for 10 min followed by overnight stirring. Acetone was laser beam wavelength of 487.99 nm. A proper con-
evaporated under a constant nitrogen flow at ~80 °C. centration of sample was loaded on the conductive
The dried samples were calcined in air at 500 °C for carbon tape for analysis.
4 h and then reduced in a H2 gas (99.995%, AGA) flow
at 500 °C for 4 h to obtain TiO2 NFs with 1 wt.% metal 4.4 Photocatalytic reaction
loading. The catalysts decorated with 1 wt.% Pt and The activity of each catalyst was tested by performing
Pd are referred to as 1%Pt–TiO2 NFs and 1%Pd–TiO2 photocatalytic CO2 reduction experiments under UV
NFs, respectively. light. Typically, 100–140 mg catalyst was dispersed in
Anchoring of 1 wt.% CdSe QDs (2.1–2.3 nm) onto 10 mL deionized water and filtered through 90 mm
the TiO2-based catalysts was obtained by first preci- polycarbonate membranes of 0.2 μm pore size. The
pitating commercial CdSe QDs twice with methanol. filter paper and catalyst were dried at room tem-
The CdSe QDs were then dissolved in dichloromethane perature overnight before the photocatalytic activity
and added to the TiO2-based catalysts (i.e. pristine NFs, was tested.
1%Pt–TiO2 NFs, and 1%Pd–TiO2 NFs). The catalysts The reaction chamber was set up as follows: The
were dried after evaporation of dichloromethane and catalyst on the filter paper was loaded in a stainless
further annealed at 600 °C for 30 min under N2 gas steel reaction chamber containing an inlet and outlet
(99.9%, AGA) to desorb the organic capping molecules. for gas. A circular stainless steel ring was placed on
For the TiO2 catalysts decorated with Pt and Pd, the the filter paper to hold it in place. Finally, a quartz
metal nanoparticles were first loaded and then CdSe window sealed with an O-ring was installed to allow
QDs were anchored onto them. These materials are light to fall on the catalyst. (For a schematic illustration
denoted as 1%CdSe–1%Pt–TiO2 NFs and 1%CdSe– and photograph of the reactor setup, see the ESM,
1%Pd–TiO2 NFs, respectively, while TiO2 NFs anchored Figs. S3(a) and S3(b).)
with CdSe QDs are labeled as 1%CdSe–TiO2 NFs. CO2 gas was passed through a bubbler containing
4.3 Structural characterization 60 mL MilliQ water (maintained at 25 °C) before
entering the reactor (maintained at 30 °C). The outlet of
The actual amount of Pd, Pt, Cd, and Se loaded in the the reactor was connected to a cold trap, maintained
catalysts was quantified on a Perkin Elmer Optima at −6 °C, which was further connected to a micro gas
5300 DV inductively coupled plasma-optical emission chromatograph equipped with a thermal conductivity
spectrometer, with which either an axial or radial detector (TCD). Before assembling the reactor, CO2
viewing mode of the plasma is possible. Each sample was first bubbled through the bubbler at a rate of
was digested using an appropriate procedure (see the 10 mL·min−1 overnight for saturation. Once the filter
ESM). The surface chemical composition was studied paper containing the catalyst was fixed in the reaction
using XPS (Kratos Axis Ultra DLD electron spectro- chamber, the reactor was assembled. The reactor was
meter, monochromated Al K source operated at closed tightly and flushed with CO2-saturated water
150 W, charge neutralizer, analyzing area 0.3 mm × for 1 h at a rate of 50 mL·min−1 to equilibrate the entire
0.7 mm). reactor. Then, the flow was reduced to 5 mL·min−1
The structures of the materials were characterized and allowed to stabilize for 30 min before irradiation.
using TEM (EFTEM, Leo 912 Omega, 120 kV, LaB6 Throughout the reaction, the flow was maintained
filament), HRTEM (JEM-2100F at 200 keV), STEM at 5 mL·min−1. Four 6W UV-B lamps (Sankyo Denki
(JEM-2100F at 200 keV in STEM mode with angular G6T5E, Japan, FWHM = 40 nm (λmax = 306 nm))
dark-field detector), and FESEM (Zeiss ULTRA plus corresponding to a total intensity of ~2 mW·cm−2 were
with an acceleration voltage of 5 kV). Raman spectra used for irradiation. The reaction was monitored for
were recorded over the range 50–1,200 cm–1 on a 5 h. Every hour a gas sample was injected into the
Horiba Jobin Yvon LabRAM HR800 equipped with a gas chromatograph equipped with TCD. The liquid
confocal Raman BX41 microscope with an excitation product condensed in the cold trap was analyzed

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


10 Nano Res.

after 5 h by a gas chromatograph (Agilent) equipped were allowed to move until all components of the
with a flame ionization detector (FID). We observed ionic forces were less than 0.03 eV·Å−1. The ΔEads of
that beyond 5 h, the activity of the catalyst gradually adsorbates is defined as ΔEads = Eadsorbate/TiO2 – Eadsorbate –
declined. ETiO2, and the ΔER of intermediates was computed as
the energy difference of each elementary reaction
4.5 Theoretical methods during CO2 reduction. The description for rutile TiO2
systems can be found in the ESM.
Ab initio calculations were performed by density
functional theory (DFT) [58] using the generalized
gradient approximation and the model of Perdew, Acknowledgements
Burke, and Ernzerhof [59] as the exchange-correlation
term. The electronic structure was solved using The theoretical simulations were performed on resour-
Vanderbilt ultrasoft pseudopotentials [60]. The kinetic ces provided by the Swedish National Infrastructure
energy cutoffs for wave functions were set to 50 and for Computing (SNIC) at the High Performance
500 Ry for the charge density. A Marzari–Vanderbilt Computing Center North (HPC2N). The Authors also
[61] smearing of 0.02 Ry was used to aid convergence. acknowledge Kempe Foundations, the Bio4Enenrgy
programme and the Artificial Leaf project under
Integration of the Brillouin zone was carried out
auspices of Knut & Alice Wallenberg Foundation for
using an 8 × 8 × 1 Monkhorst–Pack grid [62]. The DFT
funding provided.
computations were performed using the Quantum
Espresso (QE) code [63].
Electronic Supplementary Material: Supplementary
Non-spin polarized calculations were performed to
material (characterization, TEM images and Raman
describe the CO2 reduction reaction over two different
spectra of modified TiO2 samples; schematic illustration
TiO2 phases (anatase and rutile). The bulk lattice
and a photograph of the reactor setup; theoretical
parameters of anatase TiO2 (space group I41/amd)
description of rutile TiO2 system) is available in the
obtained by DFT as a = b = 3.793 Å and c = 9.647 Å
online version of this article at http://dx.doi.org/
were in good agreement with the experimental data
10.1007/s12274-016-1087-9.
[64]. For rutile TiO2, the obtained lattice parameters
were equal to a = b = 4.5922 Å and c = 2.9574 Å; these
values are also consistent with the experimental lattice References
parameters [65]. The resulting electronic band gaps
[1] Metz, B.; Davidson, O.; De Coninck, H.; Loos, M.; Meyer,
for the anatase and rutile TiO2 phases were 1.94
L. Carbon Dioxide Capture and Storage; Cambridge
and 1.85 eV, respectively, consistent with previously University Press: Cambridge, UK, 2005.
reported values (1.88 eV) using similar DFT methods [2] European Parliament Legislative Resolution of 17 December
[66]. The electronic properties and lattice parameters 2008 on the Proposal for a Directive of the European
of both bulk anatase and rutile TiO2 were obtained Parliament and the Council on the Promotion of the Use of
using an 8 × 8 × 8 Monkhorst–Pack grid for integrating Energy from Renewable Source (COM(2008)0019-C6-0046/
the Brillouin zone. 2008-2008/0016(COD)); European parliament: Strasbourg,
An anatase TiO2 (101) surface with a (2 × 1) super France, 2008.
cell containing 6 TiO2 slabs was used as the catalytic [3] Yuan, L.; Xu, Y.-J. Photocatalytic conversion of CO2 into
value-added and renewable fuels. Appl. Surf. Sci. 2015, 342,
surface. The chemical reaction on anatase TiO2 (101)
154–167.
was initiated by adsorption of a CO2 molecule on
[4] Roy, S. C.; Varghese, O. K.; Paulose, M.; Grimes, C. A.
a 5-fold coordinate Ti atom (Ti5f), followed by the Toward solar fuels: Photocatalytic conversion of carbon
consecutive addition of diverse photon electron (H+ + dioxide to hydrocarbons. ACS Nano 2010, 4, 1259–1278.
e−) pairs. During geometric optimization, the bottom [5] Liu, G. H.; Hoivik, N.; Wang, K. Y.; Jakobsen, H.
three layers were kept frozen at their bulk optimized Engineering TiO2 nanomaterials for CO2 conversion/solar
positions, while the top three slabs and adsorbents fuels. Sol. Energy Mater. Sol. Cells 2012, 105, 53–68.

| www.editorialmanager.com/nare/default.asp
Nano Res. 11

[6] Wang, C. J.; Thompson, R. L.; Baltrus, J.; Matranga, C. [19] Lo, C. C.; Hung, C. H.; Yuan, C. S.; Wu, J. F. Photoreduction
Visible light photoreduction of CO2 using CdSe/Pt/TiO2 of carbon dioxide with H2 and H2O over TiO2 and ZrO2 in a
heterostructured catalysts. J. Phys. Chem. Lett. 2010, 1, circulated photocatalytic reactor. Sol. Energy Mater. Sol.
48–53. Cells 2007, 91, 1765–1774.
[7] Dhakshinamoorthy, A.; Navalon, S.; Corma, A.; Garcia, H. [20] Dey, G. R.; Belapurkar, A. D.; Kishore, K. Photo-catalytic
Photocatalytic CO2 reduction by TiO2 and related titanium reduction of carbon dioxide to methane using TiO2 as
containing solids. Energy Environ. Sci. 2012, 5, 9217–9233. suspension in water. J. Photochem. Photobiol. A: Chem.
[8] Kubacka, A.; Fernández-García, M.; Colón, G. Advanced 2004, 163, 503–508.
nanoarchitectures for solar photocatalytic applications. [21] Guan, G. Q.; Kida, T.; Harada, T.; Isayama, M.; Yoshida, A.
Chem. Rev. 2012, 112, 1555–1614. Photoreduction of carbon dioxide with water over K2Ti6O13
[9] Kwak, B. S.; Kang, M. Photocatalytic reduction of CO2 with photocatalyst combined with Cu/ZnO catalyst under con-
H2O using perovskite CaxTiyO3. Appl. Surf. Sci. 2015, 337, centrated sunlight. Appl. Catal. A: Gen. 2003, 249, 11–18.
138–144. [22] Guan, G. Q.; Kida, T.; Yoshida, A. Reduction of carbon
[10] Li, X.; Wen, J. Q.; Low, J. X.; Fang, Y. P.; Yu, J. G. Design dioxide with water under concentrated sunlight using photo-
and fabrication of semiconductor photocatalyst for photo- catalyst combined with Fe-based catalyst. Appl. Catal. B:
catalytic reduction of CO2 to solar fuel. Sci. China Mater. Environ. 2003, 41, 387–396.
2014, 57, 70–100. [23] Yahaya, A. H.; Gondal, M. A.; Hameed, A. Selective laser
[11] Ashley, M.; Magiera, C.; Ramidi, P.; Blackburn, G.; Scott, enhanced photocatalytic conversion of CO2 into methanol.
T. G.; Gupta, R.; Wilson, K.; Ghosh, A.; Biswas, A. Chem. Phys. Lett. 2004, 400, 206–212.
Nanomaterials and processes for carbon capture and [24] Chen, H.-C.; Chou, H.-C.; Wu, J. C. S.; Lin, H.-Y. Sol–gel
conversion into useful by-products for a sustainable energy prepared InTaO4 and its photocatalytic characteristics. J.
future. Greenh. Gases: Sci. Technol. 2012, 2, 419–444. Mater. Res. 2008, 23, 1364–1370.
[12] Kumar, S. G.; Devi, L. G. Review on modified TiO2 [25] Ozcan, O.; Yukruk, F.; Akkaya, E. U.; Uner, D. Dye sensitized
photocatalysis under UV/visible light: Selected results and artificial photosynthesis in the gas phase over thin and thick
related mechanisms on interfacial charge carrier transfer TiO2 films under UV and visible light irradiation. Appl.
dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. Catal. B: Environ. 2007, 71, 291–297.
[13] Ola, O.; Maroto-Valer, M. M. Review of material design and [26] Ozcan, O.; Yukruk, F.; Akkaya, E. U.; Uner, D. Dye
reactor engineering on TiO2 photocatalysis for CO2 reduction. sensitized CO2 reduction over pure and platinized TiO2.
J. Photochem. Photobiol. C: Photochem. Rev. 2015, 24, 16–42. Top. Catal. 2007, 44, 523–528.
[14] Yu, J. G.; Low, J. X.; Xiao, W.; Zhou, P.; Jaroniec, M. [27] Nguyen, T.-V.; Wu, J. C. S.; Chiou, C.-H. Photoreduction
Enhanced photocatalytic CO2-reduction activity of anatase of CO2 over ruthenium dye-sensitized TiO2-based catalysts
TiO2 by coexposed {001} and {101} facets. J. Am. Chem. under concentrated natural sunlight. Catal. Commun. 2008,
Soc. 2014, 136, 8839–8842. 9, 2073–2076.
[15] Yamashita, H.; Fujii, Y.; Ichihashi, Y.; Zhang, S. G.; Ikeue, [28] Pathak, P.; Meziani, M. J.; Castillo, L.; Sun, Y. P. Metal-coated
K.; Park, D. R.; Koyano, K.; Tatsumi, T.; Anpo, M. Selective nanoscale TiO2 catalysts for enhanced CO2 photoreduction.
formation of CH3OH in the photocatalytic reduction of CO2 Green Chem. 2005, 7, 667–670.
with H2O on titanium oxides highly dispersed within zeolites [29] Sasirekha, N.; Basha, S. J. S; Shanthi, K. Photocatalytic
and mesoporous molecular sieves. Catal. Today 1998, 45, performance of Ru doped anatase mounted on silica for
221–227. reduction of carbon dioxide. Appl. Catal. B: Environ. 2006,
[16] Anpo, M.; Yamashita, H.; Ichihashi, Y.; Ehara, S. Photo- 62, 169–180.
catalytic reduction of CO2 with H2O on various titanium [30] Zhao, Z. H.; Fan, J. M.; Wang, Z. Z. Photocatalytic CO2
oxide catalysts. J. Electroanal. Chem. 1995, 396, 21–26. reduction using sol–gel derived titania-supported zinc-
[17] Pathak, P.; Meziani, M. J.; Li, Y.; Cureton, L. T.; Sun, Y. P. phthalocyanine. J. Clean. Prod. 2007, 15, 1894–1897.
Improving photoreduction of CO2 with homogeneously [31] Sarkar, A.; Shchukarev, A.; Leino, A.-R.; Kordas, K.;
dispersed nanoscale TiO2 catalysts. Chem. Commun. 2004, Mikkola, J.-P.; Petrov, P. O.; Tuchina, E. S.; Popov, A. P.;
1234–1235. Darvin, M. E.; Meinke, M. C. et al. Photocatalytic activity
[18] Tseng, I. H.; Chang, W.-C.; Wu, J. C. S. Photoreduction of TiO2 nanoparticles: Effect of thermal annealing under
of CO2 using sol–gel derived titania and titania-supported various gaseous atmospheres. Nanotechnology 2012, 23,
copper catalysts. Appl. Catal. B: Environ. 2002, 37, 37–48. 475711.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


12 Nano Res.

[32] Zhai, Q. G.; Xie, S. J.; Fan, W. Q.; Zhang, Q. H.; Wang, Y.; Surf. Sci. 2013, 277, 105–110.
Deng, W. P.; Wang, Y. Photocatalytic conversion of carbon [45] Roy, S.; Hegde, M. S.; Ravishankar, N.; Madras, G. Creation
dioxide with water into methane: Platinum and copper(I) of redox adsorption sites by Pd2+ ion substitution in
oxide Co-catalysts with a core–shell structure. Angew. Chem., nanoTiO2 for high photocatalytic activity of CO oxidation,
Int. Ed. 2013, 52, 5776–5779. NO reduction, and NO decomposition. J. Phys. Chem. C
[33] Liu, Y.; Zhou, S.; Li, J. M.; Wang, Y. J.; Jiang, G. Y.; Zhao, 2007, 111, 8153–8160.
Z.; Liu, B.; Gong, X. Q.; Duan, A. J.; Liu, J. et al. [46] Galian, R. E.; De la Guardia, M.; Pérez-Prieto, J. Size reduction
Photocatalytic reduction of CO2 with water vapor on surface of CdSe/ZnS core−shell quantum dots photosensitized by
La-modified TiO2 nanoparticles with enhanced CH4 selectivity. benzophenone: Where does the Cd(0) go? Langmuir 2011,
Appl. Catal. B: Environ. 2015, 168–169, 125–131. 27, 1942–1945.
[34] Li, Q. Y.; Zong, L. L.; Li, C.; Yang, J. J. Photocatalytic [47] Trinh, T. T.; Mott, D.; Thanh, N. T. K.; Maenosono, S.
reduction of CO2 on MgO/TiO2 nanotube films. Appl. Surf. One-pot synthesis and characterization of well defined
Sci. 2014, 314, 458–463. core–shell structure of FePt@CdSe nanoparticles. RSC Adv.
[35] Long, M. C.; Cai, W. M.; Cai, J.; Zhou, B. X.; Chai, X. Y.; 2011, 1, 100–108.
Wu, Y. H. Efficient photocatalytic degradation of phenol [48] Yasuo, I. Recent advances in the photocatalytic conversion
over Co3O4/BiVO4 composite under visible light irradiation. of carbon dioxide to fuels with water and/or hydrogen using
J. Phys. Chem. B 2006, 110, 20211–20216. solar energy and beyond. Coord. Chem. Rev. 2013, 257,
[36] Tada, H.; Mitsui, T.; Kiyonaga, T.; Akita, T.; Tanaka, K. 171–186.
All-solid-state Z-scheme in CdS–Au–TiO2 three-component [49] Wu, M.-C.; Sápi, A.; Avila, A.; Szabó, M.; Hiltunen, J.;
nanojunction system. Nat. Mater. 2006, 5, 782–786. Huuhtanen, M.; Tóth, G.; Kukovecz, Á.; Kónya, Z.; Keiski,
[37] Yang, C. Y.; Wang, W. D.; Shan, Z. C.; Huang, F. Q. R. et al. Enhanced photocatalytic activity of TiO2 nanofibers
Preparation and photocatalytic activity of high-efficiency and their flexible composite films: Decomposition of organic
visible-light-responsive photocatalyst SnSx/TiO2. J. Solid dyes and efficient H2 generation from ethanol–water mixtures.
State Chem. 2009, 182, 807–812. Nano Res. 2011, 4, 360–369.
[38] Robel, I.; Kuno, M.; Kamat, P. V. Size-dependent electron [50] Kamat, P. V. Meeting the clean energy demand: Nanostructure
injection from excited CdSe quantum dots into TiO2 nano- architectures for solar energy conversion. J. Phys. Chem. C
particles. J. Am. Chem. Soc. 2007, 129, 4136–4137. 2007, 111, 2834–2860.
[39] Zhang, Q.-H.; Han, W.-D.; Hong, Y.-J.; Yu, J.-G. [51] Ishitani, O.; Inoue, C.; Suzuki, Y.; Ibusuki, T. Photocatalytic
Photocatalytic reduction of CO2 with H2O on Pt-loaded reduction of carbon dioxide to methane and acetic acid by an
TiO2 catalyst. Catal. Today 2009, 148, 335–340. aqueous suspension of metal-deposited TiO2. J. Photochem.
[40] Koči, K.; Matějů, K.; Obalová, L.; Krejčíková, S.; Lacný, Z.; Photobiol. A: Chem. 1993, 72, 269–271.
Plachá, D.; Čapek, L.; Hospodková, A.; Šolcová, O. Effect [52] Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C. F.;
of silver doping on the TiO2 for photocatalytic reduction of Hummelshøj, J. S.; Dahl, S.; Chorkendorff, I.; Nørskov, J. K.
CO2. Appl. Catal. B: Environ. 2010, 96, 239–244. Discovery of a Ni-Ga catalyst for carbon dioxide reduction
[41] Li, X. K.; Zhuang, Z. J.; Li, W.; Pan, H. Q. Photocatalytic to methanol. Nat. Chem. 2014, 6, 320–324.
reduction of CO2 over noble metal-loaded and nitrogen-doped [53] Grabow, L. C.; Mavrikakis, M. Mechanism of methanol
mesoporous TiO2. Appl. Catal. A: Gen. 2012, 429–430, synthesis on Cu through CO2 and CO hydrogenation. ACS
31–38. Catal. 2011, 1, 365–384.
[42] Liu, L. J.; Li, Y. Understanding the reaction mechanism of [54] Sorescu, D. C.; Al-Saidi, W. A.; Jordan, K. D. CO2 adsorption
photocatalytic reduction of CO2 with H2O on TiO2-based on TiO2(101) anatase: A dispersion-corrected density
photocatalysts: A review. Aerosol Air Qual. Res. 2014, 14, functional theory study. J. Chem. Phys. 2011, 135, 124701.
453–469. [55] He, H. Y.; Zapol, P.; Curtiss, L. A. A theoretical study of
[43] Meng, X. Q.; Ouyang, S. X.; Kako, T.; Li, P.; Yu, Q.; CO2 anions on anatase (101) surface. J. Phys. Chem. C
Wang, T.; Ye, J. H. Photocatalytic CO2 conversion over alkali 2010, 114, 21474–21481.
modified TiO2 without loading noble metal cocatalyst. Chem. [56] Valdés, Á.; Qu, Z.-W.; Kroes, G.-J.; Rossmeisl, J.; Nørskov,
Commun. 2014, 50, 11517–11519. J. K. Oxidation and photo-oxidation of water on TiO2
[44] Kong, D.; Tan, J. Z. Y.; Yang, F.; Zeng, J. L.; Zhang, X. W. surface. J. Phys. Chem. C 2008, 112, 9872–9879.
Electrodeposited Ag nanoparticles on TiO2 nanorods for [57] Chen, H. H.; Park, H. Millis, A. J.; Marianetti, C. A. Charge
enhanced UV visible light photoreduction CO2 to CH4. Appl. transfer across transition-metal oxide interfaces: Emergent

| www.editorialmanager.com/nare/default.asp
Nano Res. 13

conductance and electronic structure. Phys. Rev. B 2014, 90, zone integrations. Phys. Rev. B 1976, 13, 5188–5192.
245138. [63] Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.;
[58] Kohn, W.; Sham, L. J. Self-consistent equations including Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni,
exchange and correlation effects. Phys. Rev. 1965, 140, M.; Dabo, I. et al. Quantum espresso: A modular and open-
A1133. source software project for quantum simulations of materials.
[59] Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized J. Phys.: Condens. Mat. 2009, 21, 395502.
gradient approximation made simple. Phys. Rev. Lett. 1996, [64] Howard, C. J.; Sabine, T. M.; Dickson, F. Structural and
77, 3865–3868. thermal parameters for rutile and anatase. Acta Cryst. B
[60] Vanderbilt, D. Soft self-consistent pseudopotentials in a 1991, 47, 462–468.
generalized eigenvalue formalism. Phys. Rev. B 1990, 41, [65] Swope, R. J.; Smyth, J. R.; Larson, A. C. H in rutile-type
7892–7895. compounds: I. Single-crystal neutron and X-ray diffraction
[61] Marzari, N.; Vanderbilt, D.; De Vita, A.; Payne, M. C. study of H in rutile. Amer. Mineral. 1995, 80, 448–453.
Thermal contraction and disordering of the Al (110) surface. [66] Landmann, M.; Rauls, E.; Schmidt, W. G. The electronic
Phys. Rev. Lett. 1999, 82, 3296–3299. structure and optical response of rutile, anatase and brookite
[62] Monkhorst, H. J.; Pack, J. D. Special points for brillouin- TiO2. J. Phys.: Condens. Mat. 2012, 24, 195503.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

Potrebbero piacerti anche