You are on page 1of 8

Applied Catalysis B: Environmental 298 (2021) 120635

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Enhanced CH4 yields by interfacial heating-induced hot water steam during


photocatalytic CO2 reduction
Xiantao Hu 1, Zhanjun Xie 1, Qian Tang 1, Heng Wang, Lianbin Zhang *, Jingyu Wang *
Key Laboratory of Material Chemistry for Energy Conversion and Storage (Ministry of Education), School of Chemistry and Chemical Engineering, Huazhong University
of Science and Technology, Wuhan, 430074, Hubei, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: The adsorption and activation of H2O vapor on photocatalyst is highly important for CO2 reduction yet having
Photocatalytic CO2 reduction been overlooked among the large amounts of reports. Herein, we demonstrate a preferential H2O adsorption
H2O adsorption mechanism to replace the previously assumed competitive adsorption, that is, H2O molecules were pre-adsorbed
Photothermal membrane
on the catalytic sites and dissociated to provide protons for the activation of the adsorbed CO2 molecules. Thanks
Water steam
Solid-gas reaction
to a photothermal membrane (PM)-induced interfacial heating, the generation of hot steam dramatically in­
creases the partial pressure of water vapor from 17 to 43 kPa at ambient pressure, which favors H2O adsorption
on TiO2 surface. The well-designed TiO2+PM system achieves selective reduction of CO2 to CH4 with a high
evolution rate of 18.3 μmol g− 1 h− 1, about 3.3 times of that in the absence of PM. The light-to-heat conversion
makes the utilization of light energy more efficient, not merely for the photogeneration of charge carriers.

1. Introduction gas-phase conditions is more attractive because it favors the high


selectivity to hydrocarbon production and the reagents are completely
Nowadays, the increasing CO2 emission in the atmosphere by the green and rich in nature [15–18]. Meanwhile the merits of stability and
depletion of carbon-rich fossil fuels has caused serious social issues recyclability of the solid-gas catalysis make it applicable in industrial
including the destruction of ecological balance and the shortage of production if the main obstacle, i.e. relatively low conversion efficiency,
carbonaceous resources. Inspired by natural photosynthesis, the pho­ can be overcome. To tackle this issue, numerous efforts have been
tocatalytic technique has been explored as a promising strategy to focused on the improvement of light absorption and charge carriers
address the issues by utilizing abundant solar energy to convert CO2 into separation of photocatalysts [19–26]. Most recently, the CO2 adsorption
valuable fuels [1–6]. Among the existing photocatalytic systems, two on the catalyst surface has been gradually realized as the crucial step in
approaches, i.e., gas-phase and liquid-phase reactions, have been the whole reaction process [27–30]. There are increasing reports on the
developed for CO2 conversion. In the liquid-phase reaction, metal combination of photocatalysts with microporous materials due to their
complexes and metal–organic frameworks are usually employed as the superior specific surface area and CO2 uptake ability, leading to the
photocatalysts which realized high catalytic reactivity towards CO2 enhanced CO2 conversion efficiency [31–35].
reduction [7–9]. However, the reduction can only occur in the anhy­ As another reactant molecule, the H2O vapor also plays an important
drous or less water system containing organic sacrificial reagent to role in the CO2 reduction reaction due to its contribution to the oxida­
overcome the sluggish kinetics of oxidation process, especially for the tion cycle. Even its presence is known to be essential, the impact of H2O
fact that CO is the main product and hydrocarbon fuel generation is vapor on the efficiency of CO2 reduction has always been overlooked in
quite limited [10–12]. Moreover, the presence of carbon-containing comparison to the large amounts of works related to the development of
sacrificial donors brought uncertainties in quantifying the reduction various photocatalysts [13]. Actually, it has been proved that the
yield since they could also contribute to the reduction products or dissociation of H2O on the catalyst surface is the first step of CO2 acti­
participate in the reduction process in an indirect way [13,14]. vation by providing protons to react with the surface adsorbed CO2
In contrast, the photocatalytic CO2 reduction with H2O vapor under molecules [36–38]. That is, the CO2 reduction efficiency is normally

* Corresponding authors.
E-mail addresses: zhanglianbin@hust.edu.cn (L. Zhang), wangjingyu@hust.edu.cn (J. Wang).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.apcatb.2021.120635
Received 3 June 2021; Received in revised form 5 August 2021; Accepted 14 August 2021
Available online 18 August 2021
0926-3373/© 2021 Elsevier B.V. All rights reserved.
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

restricted by the kinetically sluggish water oxidation cycle [39]. As water (90 mL) under magnetic agitation. When the mixture was
discovered in our previous study, the photocatalytic activity of the completely added, the resulted suspension was heated to 70 ◦ C and
whole process strongly depended on the content of H2O vapor and CH4 stirred mechanically for about 2 h until it was viscous. Then 1 M NaOH
evolution rate obviously increased with the humidity [40]. In addition, (300 mL) was added into the condensed suspension and was stirred at 70
the comparison of competitive adsorption demonstrated that the ◦
C under mechanical stirring for 12 h. LPT was obtained and washed
pre-adsorption of water did not affect the CO2 adsorption ability or the repeatedly by deionized water. LPT precursor (9 g) were dispersed into
photocatalytic efficiency [27,39]. It can be deduced that the overall isopropyl alcohol (6 mL) and stirred for 40 min. Next, fluoric acid (0.5
reaction will be accelerated if increasing the H2O coverage with mL) was added into the above suspension to synthesizeTiO2 photo­
co-adsorption of CO2 on the photocatalyst. catalyst after 12 h of solvethermal treatment at 180 ◦ C.
Aiming at increasing H2O coverage and reaction kinetics in the CO2
reduction process, herein we employed the photothermal membrane 2.3. Characterization
(PM) as the first example of combination with photocatalytic CO2 con­
version, through which the interfacial heating-produced hot water The morphology was obtained on microscopes including field
steam dramatically promoted the reduction of CO2. The scheme of PM- emission scanning electron microscopy (FESEM, Hitachi SU8010) and
assisted TiO2 photocatalysis (TiO2+PM) for CO2 reduction is illustrated transmission electron microscopy (TEM, Tecnai G2 F30). Fourier-
in Fig. 1. PM was completely wetted and adhered to the reactor wall to transform infrared (FT-IR) spectra were monitored on a VERTEX 70.
realize the spatial isolation from TiO2 photocatalyst. In this way, it can The Pyris1 TGA (PerkinElmer Instruments) was used to test the ther­
receive the scattered light and then generate hot water steam in the mogravimetric curves of samples. The infrared camera (MAG32,
system by the photothermal effect. Meanwhile, the location of PM Shanghai Magnity Electronics CO., Ltd) was settled to monitor the
avoids the light shielding and direct heating of TiO2 surface. The PM- temperature on the photocatalyst surface. The measurements of contact
assisted photocatalytic strategy can effectively accelerate H2O evapo­ angle were performed on an OCA 15EC system. X-ray powder diffraction
ration to increase H2O coverage on the catalyst surface without extra (XRD, SmartLab-SE) was used to record the crystal structures of as-
energy consumption and therefore enhance the efficiency and selectivity prepared samples. UV–vis diffuse reflectance spectrum (DRS) was con­
for CH4 yields. ducted on a UV–vis spectrophotometer (UV-3600, Shimadzu, Japan)
with BaSO4 power as a reflectance standard. N2 adsorption and
2. Experimental desorption isotherms were examined using an ASAP 2020 (Micro­
meritics Instruments, USA). Electrochemical impedance spectra (EIS)
2.1. Synthesis of polypyrrole coated Airlaid paper (PPy/Airlaid paper) under different temperature was carried out in a 0.1 M electrolyte so­
lution containing KCl, K3Fe(CN)6 and K4Fe(CN)6. The photocurrent re­
1.13 μL of FeCl3⋅6H2O aqueous solution was added into the pyrrole sponses of as-prepared samples under different temperature were
aqueous solution, and then the airlaid paper was placed on the solution measured in a 0.1 M Na2SO4 solution using the electrochemical work­
surface. With the initiation of Fe3+, the pyrrole monomers polymerized station (CHI650E, Chenhua) equipped with a three-electrode system and
on the Airlaid paper surface to form the polypyrrole coated Airlaid constant temperature water bath.
paper. After reacting for 30 min, the PPy/Airlaid paper was separated
from the solution and washed several times with ethanol. Finally, the 2.4. Photocatalytic test for CO2 conversion
resultant PPy/Airlaid paper was dried in an oven for 5 h under 40 ◦ C.
Photocatalytic performance toward CO2 conversion over the syn­
thesized catalysts was tested in a sealed batch reactor under mild con­
2.2. Synthesis of TiO2 photocatalyst
dition. The volume of such a sealed container for photocatalytic reaction
is 350 mL. The customized photocatalytic reactor was equipped with
TiO2 photocatalyst was synthesized using the lamellar protonated
water jacket and circulating water pump. Thus the temperature of the
titanate (LPT) as a precursor. The mixture containing tetrabutyl titanate
water vapor in the reactor can be adjusted by the circulating water in­
(9 mL) and absolute ethanol (66 mL) was slowly added into deionized
side the water jacket. A 300 W Xe lamp (PLS-SXE300D, Beijing Per­
fectlight, China) was used as the light source. The full spectrum of the Xe
lamp is shown in Fig. S1, showing the range of main wavelength as
325~780 nm. Before reaction, 50 mg of catalyst powders were placed on
the glass dish that positioned 5 cm of distance under light source. PPy/
Airlaid paper was tailored to a square with a side length of 4.5 cm and
exposed area of 20.25 cm2, which was placed vertically inside the sealed
reactor. The average light intensity, calculated from the measured
values of four corners of PPy/Airlaid paper, was used to quantify the
light intensity arriving at the paper which could be adjusted by the
power of light source. PPy/Airlaid paper were positioned 3 cm of dis­
tance below the glass dish. The deionized water (2 mL) was injected into
the PPy/Airlaid paper. Before the light irradiation, the photocatalytic
reactor was thoroughly vacuum-treated, then CO2 gas was introduced
into the circulation system. The gas products from the reduction reaction
were analyzed by gas chromatography (GC-2014C, Shimadzu Corp.
Japan) equipped with thermal conductivity detector (TCD) for H2
detection and flame ionization detector (FID) for CO and CH4 detections.
To eliminate the oxygen interference by the air, the O2 product from the
oxidation reaction was monitored by an optic fiber oxygen sensor
(Ocean-Optics), which was sealed in the photocatalytic reactor to be
isolated from the surrounding environment. The cycling test of photo­
Fig. 1. Schematic apparatus for photocatalytic CO2 conversion assisted by catalytic CO2 conversion was conducted by recording the gas evolution
photothermal membrane. rates of five cycles under identical conditions. After each cycle, TiO2

2
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

photocatalyst was dried in a vacuum oven to remove the surface water steam production. The steam generation performance of the
adsorbed water and PPy/Airlaid paper was wet again by the deionized photothermal membrane was also evaluated by recording
water. time-dependent mass change and the corresponding evaporation rate
under solar radiation (Figs. S5 and 2 e). Under 2.0 kW m-2 of light
3. Results and discussion illumination, the evaporation efficiency reached 60.1 %, suggesting the
excellent steam generation performance of PPy/Airlaid paper [43–46].
3.1. Characterization of morphology and structure The cycling test confirmed that the PPy membrane well maintained the
evaporation performance within five repeated cycles (Fig. S6), which
The membrane containing photothermal material is considered to be suggested its applicable to the long-term reaction.
an effective technique for solar-driven hot steam generation. Poly­ Having confirmed the interfacial heating-induced water steam pro­
pyrrole (PPy) as a promising conductive polymer shows wide light ab­ duction on PPy membrane, we used it to provide more H2O vapor for
sorption from ultraviolet to near infrared light and an excellent participating in the photocatalytic reduction of CO2 over TiO2 catalysts.
photothermal transduction [41]. PPy decorated Airlaid paper as a TiO2 photocatalysts were prepared from the lamellar protonated tita­
typical PM material was introduced into the photocatalytic system. The nate precursor according to our previous report [27,47]. The crystal
PPy membrane, synthesized by in situ polymerization of pyrrole on the structure, morphology, and surface properties were investigated by
Airlaid paper, presents a rough outmost surface by the uniform deco­ XRD, TEM, DRS and N2 adsorption–desorption measurements. All the
ration of PPy nanoparticles, relative to a rather smooth surface of the diffraction peaks in Fig. 3a can be assigned to the anatase phase of TiO2
paper substrate (Fig. 2a, b). Fourier-transform infrared (FT-IR) spectra crystals, according to the standard pattern (65-5714). TEM and
of both Airlaid paper and PPy/Airlaid paper display the characteristic high-resolution TEM (HRTEM) observations in Fig. 3b and c present the
absorption bands of O–H, C–H, C– – O, C–O, and C–C groups due to well-defined corners and edges of TiO2 catalysts. The crystal lattice
the existence of cellulose (Fig. S2) [42]. Differently, two absorption spacing of 0.352 and 0.235 nm corresponds to the {101} and {001}
peaks of C–N (923 cm− 1) and C– – C in pyrrole ring (1555 cm− 1) facets and the interfacial angle of them is 68.3 ◦ (Fig. 3d), which
emerged in PPy/Airlaid paper [41], both of which confirmed the coating confirmed the formation of nanosized anatase TiO2 crystals with coex­
of paper surface with PPy membrane. As a photothermal membrane, the posed {101} and {001} facets. DRS in Fig. S7 displays that the light
thermal stability of the PPy/Airlaid paper should be considered in terms absorption edge of TiO2 photocatalysts extends up to 389 nm, within the
of the durability and recyclability of the photocatalytic system. The light wavelength of Xe lamp. The N2 adsorption–desorption isotherm of
thermogravimetric (TG) analysis shows the good stability of pristine TiO2 photocatalyst is typical type IV, which provides a high value of
paper substrate with or without PPy membrane at the temperature lower specific surface area as 118 m2 g− 1 consisting of a narrow mesopore size
than 250 ◦ C in air atmosphere (Fig. S3). The temperature at the surface distribution of 4.86 nm (Fig. S8).
of dry PPy/Airlaid paper was measured to be 76~113 ◦ C under light
illumination with intensity varying from 1.0 to 2.5 kW m-2 (Fig. 2c, d). 3.2. Evaluation of photocatalytic activity and stability
Thus the structure stability of the photothermal membrane can be un­
changed under normal operating conditions. The surface hydrophilicity Photocatalytic CO2 reduction was carried out in a solid-gas catalytic
of the material is considered to be an important factor in the interfacial system under 2.0 kW m− 2 of light illumination using H2O vapor as
heating-induced hot water steam production. The water contact angles electron donors. In the photocatalytic system, the main products, CH4
of PPy/Airlaid paper and Airlaid paper reach 0 ◦ (Fig. S4). The result and CO, were obtained through the 8-electron and 2-electron reduction
indicates the excellent hydrophilicity of PPy membrane after coating pathways, respectively. The signal of H2 product in the gas chroma­
Airlaid paper, which ensures the sufficient water supply during the tography spectrum was not detectable (Fig. S9), which indicated that the

Fig. 2. Scanning electron microscopy (SEM) images of (a) Airlaid paper and (b) PPy/Airlaid paper. (c) Changes of temperature at the surface of PPy/Airlaid paper
under light illumination of varied intensities. (d) Infrared images of PPy/Airlaid paper during light illumination. (e) Solar to evaporation energy conversion efficiency
and evaporation rate of different optical density.

3
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

Fig. 3. (a) XRD pattern, (b) TEM and (c, d) HRTEM images of TiO2 photocatalyst. The vertical column in (a) is the standard pattern (65-5714) of TiO2 crystal.

dissociation of H2O provided protons for CO2 reduction rather than be illumination. The results suggest that PM cannot function as catalysts for
reduced to produce H2. In comparison to a considerable amount of H2 CO2 conversion but exert a positive influence on the photocatalytic
evolution from the liquid reactants [48–52], the gas photocatalytic process. The signal of m/z = 17 in the isotopic labeling experiment can
system is beneficial to suppress the side reaction of H2O reduction and be corresponded to 13CH4 (Fig. S10), confirming the origin of the
improve the selectivity for CO2 conversion. Fig. 4a, b shows the product from CO2 conversion. The signals of m/z = 16 and 15 probably
time-course CH4 and CO evolution in the systems of PM, TiO2, and come from the fragments of 13CH4 together with those from the
TiO2+PM. Table S1 presents the average gas evolution rates and elec­ pre-adsorbed 12CO2 on the surface of catalyst. Fig. 4c displays the cyclic
tron consumption (Re). The TiO2 photocatalysts present average evolu­ conversion efficiency of TiO2+PM system during five consecutive runs.
tion rates of 5.5 μmol g-1 h-1 for CH4 and 3.6 μmol g-1 h-1 for CO, close to The negligible changes in evolution rates of products suggest the
the results of our previous reports [17,44]. Interestingly, the introduc­ excellent durability of PM material and photocatalyst for a long-term
tion of PM dramatically increased the efficiency of CO2 reduction over reaction. The structural characterizations of the recycled TiO2 cata­
TiO2 photocatalysts, especially for the selective production of CH4 lysts in Figs. S11 and S12 further demonstrate the stability of crystal
showing about 3.3 times of the original yields (18.3 μmol g-1 h-1). It is phase and chemical structure after the reaction.
noted that the CO production are extremely low (<1 μmol g-1 h-1) in the
reaction system only containing PM (no TiO2 photocatalysts). Such low
CO product may come from the slight decomposition of PM under

Fig. 4. Time-dependent evolution rate of (a) CH4 and (b) CO products in the systems of PM, TiO2, and TiO2+PM. The photocatalytic reactions were carried out in a
batch system under standard atmospheric pressure and light illumination. (c) Cycling test of photocatalytic CO2 conversion in the TiO2+PM system.

4
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

3.3. Mechanism of enhanced CH4 yield 45 ◦ C respectively. The results illustrated that the higher temperature
could facilitate the transport of photogenerated charge carriers, leading
To clarify the effect of the photothermal membrane in light-to-heat to the higher yields of effective electrons for CO2 reducion.
conversion and water steam generation, we systematically study the On the other hand, even at the same degree of TC, the k value of CH4
influence of temperature and water vapor on the CO2 conversion effi­ production is higher in the TiO2+PM system than that in the TiO2 system
ciency over TiO2 photocatalysts. In the batch reactor, the temperature of (Fig. 5d). It is noted that the PM-induced difference in TC (73 → 86 ◦ C)
gas mixture (TG) rises quickly and then reaches a steady state under light can only cause ~ 20 % of increase in CH4 evolution rate based on the
illumination. It is found that the introduction of PM leads to an obvious Arrhenius plot. Hence the dramatically improved reaction rate in
increase in TG from 56 to 78 ◦ C (Fig. S13). Accordingly, the surface TiO2+PM system suggests the non-negligible effect of ambient medium
temperature of catalysts (TC) in TiO2+PM system reached 86 ◦ C, much surrouding the catalysts. Actually, the PM-induced light-to-heat con­
higher than that in the TiO2 system (73 ◦ C, Fig. 5a–c). Since PM is version promoted the water evaporation to produce more H2O vapor in
spatially isolated from TiO2 photocatalysts in the reactor, and it cannot the reactor. Then the role of the partial pressure of H2O in the photo­
directly heat them via their physical contact but the thermal energy can catalytic CO2 reduction process should be discussed separately from the
be transferred from the PM-induced hot H2O vapor. Consistent with the temperature. Without the addition of water, the evolution rate of CH4 on
results of steam evaporation, the evolution rates of CH4 and CO TiO2 declined to 9 % of the original value (5.5 → 0.5 μmol g− 1 h− 1),
increased with the light intensity arriving at the PM surface (IP) even while the evolution rate of CO showed a slight increase to 4.4 μmol g− 1
though the light intensity at the photocatalyst surface (IC) was kept h− 1. The obvious changes in CH4 and CO evolution rates were also
unchanged (Fig. S14). The results indicated that PM-induced steam observed in TiO2+PM system (Fig. 6a, b). In this case, the electron donor
evaporation both increased TG and TC values, which might exhibit a presumably comes from the pre-adsorbed H2O on the vessel and catalyst
combined effect on the CO2 conversion. surface. Normally, the reduction of CO2 to CH4 is relatively difficult due
An apparent influence of the elevated temperature lies in increasing to the requirements of accepting eight electrons and eight protons.
the kinetic energy of CO2 and H2O molecules on the catalyst surface. Therefore, the low concentration of H2O vapor cannot provide enough
Thus the reaction can be accelerated by the existence of more activated electrons and protons, which restricted the CH4 evolution. With
molecules at the surface catalytic sites. In light of the linear increase in increasing the partial pressure of H2O vapor (pH2O), the CH4 evolution
CH4 concentration with irradiation time, its production rate constant (k) rate in TiO2 system almost increased linearly in a wide range of pH2O =
can be calculated based on a pseudo-zero order model. The temperature- 2~17 kPa (Fig. 6c). In heterogeneous catalytic system, the reaction rate
dependent kinetics experiment was conducted using circulating water is proportional to the fractional coverage of each reactant on the surface
flowing around the photocatalyst to avoid the photothermal effect of PM catalytic sites (Eq. (1)). With regard to solid-gas phase catalysis, the
on TC. The Arrhenius plot in Fig. 5d clearly demonstrated that the coverage of gaseous reactant depends on its partial pressure in gas
elevation of the surface temperature of photocatalyst favored the mixture. There are two gaseous reactants involved in the reaction, i.e.
reduction of CO2 to CH4 product. EIS in Fig. 5e display the effect of CO2 and H2O, which were reported to present a competitive adsorption
temperature on the electronic conductivity of TiO2 electrode. The radius mode with a different adsorption-desorption equilibrium constant on
of TiO2 decreased with the increase of temperature, which indicated the the catalyst surface [53,54], to produce the main products as CH4, CO,
more efficient charge transfer at higher temperature. The improved and O2. The relationship between the reaction rate and fraction of sites
interfacial charge transfer by temperature can be further confirmed by coverage can be established using Langmuir-Hinshelwood mechanism
photocurrent measurement (Fig. 5f). The photocurrent response grad­ (Eq. (2)).
ually increased when elevating the temperature from 25 ◦ C to 30 ◦ C and
r = kθCO2 θH2 O (1)

Fig. 5. (a) Monitoring of temperature on the


photocatalyst surface in the reactor after
turning on 300 W Xe lamp. Light intensity: IP =
2.0 kW m− 2 on PPy membrane surface; IC = 5.0
kW m− 2 on TiO2 catalyst surface. Infrared
camera images of TiO2 surface in (b) TiO2 and
(c) TiO2+PM systems. (d) Temperature-
dependent kinetic analysis of CH4 production
rate. TiO2 catalysts were surrounded by the
circulating water in a jacketed reactor to adjust
the temperature. (e) EIS and (f) photocurrent
response of TiO2 at 25 ◦ C, 30 ◦ C and 45 ◦ C.

5
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

Fig. 6. (a) Comparison in GC spectra of CH4 and CO products in the TiO2+PM system between with and without adding water. (b) Comparison in gas evolution rate
in the reaction systems between with and without adding water. (c) Influence of partial pressure of H2O vapor on the CH4 evolution rate over TiO2 photocatalysts
under light irradiation. (d) Plot of gas evolution rate as function of pH2O in TiO2+PM system. Solid line: changes in CH4 and CO production rates by adjusting Ip from
1.0 to 2.5 kW m− 2. Dot line: changes in CH4 and CO production rates by temperature of PM. (e) Changes in gas temperature in the reactor and (f) comparison of CH4
production rate under different conditions of water vapor. In (e, f), the orange and blue lines are the results under normal conditions of TiO2 and TiO2+M; the green
and red lines refer to the hot H2O vapor with similar temperature to that in TiO2+M system being injected into the reaction system and the difference between green
and red lines are that hot H2O vapor was continuously heating during the photocatalytic reaction or not. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article).

KH2 O pH2 O KCO2 pCO2 both CH4 and CO with gas temperature were plotted as dot line in
r=k 2
(2) Fig. 6d. Interestingly, these two kinds of lines highly overlapped each
(1 + KH2 O pH2 O + KCO2 pCO2 + KCO pCO + KCH4 pCH4 + KO2 pO2 )
other, which indicated that PM mostly functioned as the source of hot
Where r is the reaction rate, Ɵ represents the fractional coverage of H2O vapor for CO2 to CH4 conversion over photocatalysts. As further
gaseous reactant on the active sites of catalyst surface, p and K are the evidence, hot H2O vapor with similar temperature was directly injected
partial pressure and adsorption-desorption equilibrium constant for into the reaction system to replace PM. The conversion efficiency of CO2
each reactant and detectable product, respectively. The Eq. (2) can be to CH4 appears to be close to that of TiO2+PM system with similar
simplified to Eq. (3) based on the following assumptions: (a) the temperature of hot H2O vapor. Then the CH4 production rate suddenly
adsorption-desorption equilibrium constant of H2O (KH2O) is proved to declined when the temperature of H2O vapor dropped (Fig. 6e and f).
be much lower than that of CO2 (KCO2) on the TiO2 surface [53,54]; (b) In contrast to the crucial role of pH2O, the effect of pCO2 on the re­
the yield of CO is extremely low even it can be moderately adsorbed; (c) action rate appears to be limited. Under constant pH2O, the production of
the yields of CH4 and O2 are low and both of them are weakly adsorbed. CH4 and CO showed a similar trend with the increase of pCO2 in both
TiO2 and TiO2+PM systems (Fig. 7a and b). The results suggested the
r=k
KH2 O pH2 O KCO2 pCO2
(3) same influence of pCO2 on CH4 and CO production, meanwhile, the
(1 + KCO2 pCO2 )2 introduction of PM did not alter the adsorption property of CO2 mole­
cules on the TiO2 surface. At low pCO2 (≤ 2 kPa), the KCO2•pCO2 value is
According to Eq. (3), r is proportional to the partial pressure of H2O far less than 1, therefore CO2 coverage on TiO2 surface presents a linear
under constant IC, TC, and pCO2. Thus, such a simple kinetic model clearly increase with pCO2. Thus the reaction rate is directly proportional to the
demonstrates the linear relationship between CH4 evolution rate and change of pCO2 in gas mixture, which is in accordance with eq. 3.
pH2O in the TiO2 system, as shown in Fig. 6c. As above-mentioned, the However, the increase in CH4 and CO production gradually slowed
introduction of PM into the TiO2 system significantly improved the down and then reached a reaction equilibrium state at higher pCO2,
generation of the hot water. The elevated temperature accordingly lead presumably due to the entire coverage of CO2 gas molecules on the
to a higher saturated vapor pressure, i.e. higher pH2O in TiO2+PM system available catalytic active sites. Obviously, the relationship between r
than that in TiO2 system. Further, the value of pH2O in TiO2+PM system and pCO2 is inapplicable to the competitive adsorption mechanism of eq.
can be directly regulated by IP when keeping IC unchanged. Through 3. The result in the region of pCO2 = 2~85 kPa implies that the increased
adjusting the value of IP from 1.0 to 2.0 kW m− 2, the TG was measured to CO2 molecules might not occupy the adsorption sites for H2O molecules.
be 63~78 ◦ C, corresponding to the pH2O varying from 22~43 kPa. It is H2O molecules were pre-adsorbed on the catalytic sites. Both theoretical
found that the plot of both CH4 and CO evolution rates nearly shows calculations and experiment results revealed that CO2 adsorbed weakly
linearity as a function of pH2O (solid line in Fig. 6d). Among them, the on TiO2 surface but the co-adsorption with H2O could increase the CO2
CH4 production shows a much steeper slope, which suggests that more adsorption energy. As reported, water molecules were found to easily
H2O vapor favors the selective reduction of CO2 to CH4 relative to CO adsorb on the Ti sites, leading to the formation of a hydrogen bond
product. Another approach was adopted for adjusting pH2O just by between the H2O molecule and one of the O atoms of CO2 [55,56]. In
controlling the temperature of circulating water flowing around the contrast, CO2 molecule is least strongly bound to TiO2 surface in the
reactor while IP was kept unchanged. The changes in production rates of absence of water. The increased CO2 adsorption in the presence of

6
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

Fig. 7. Effect of the CO2 concentration on the CH4 and CO evolution rates in (a) TiO2 and (b) TiO2+M systems. The partial pressure of water vapor was kept constant
and N2 was used as balance gas. (c) Average rate of O2 evolution from the photocatalytic systems during 3 h of light illumination.

co-adsorbed H2O has also been confirmed in other systems, such as ZrO2, kPa at normal pressure, among which the latter contributed more to the
carbon, metal–organic framework, perovskite, et al. [57–61]. Therefore, selective reduction of CO2 to CH4. Through the temperature and
the pre-adsorbed H2O molecules on TiO2 surface can facilitate the concentration-dependent kinetic experiments, the adsorption mecha­
co-adsorption of CO2 due to the formation of H-bonding, which favored nism on the photocatalyst surface has been modified as the preferential
the proton transfer to activate the neighboring CO2 and yield the in­ H2O adsorption mode to replace the previously assumed competitive
termediates of bicarbonate species. On this premise, the value of ƟH2O adsorption between CO2 and H2O. The H2O molecules were pre-
could be considered as a constant under stable TG and pH2O, so the ki­ adsorbed on the Ti sites of TiO2 surface to the formation of H-bonding
netic model can be modified as Eq. (4), which clarified the relationship with CO2 molecules. Then the H2O molecules were dissociated to pro­
between r and pCO2. vide protons for the activation of the neighboring CO2 molecules and
yield the intermediates of bicarbonate species.With the assistance of PM-
KCO2 pCO2
r = kθH2 O (4) induced light-to-heat conversion, the light energy can be more effi­
1 + KCO2 pCO2
ciently utilized for water vapor production together with the photo­
Based on the above discussions, the conversion efficiency strongly generation of charge carriers. This work not only presents new insights
depends on the partial pressure of H2O vapor and CH4 evolution rate into understanding the roles of CO2 and H2O vapor in the solid-gas
obviously increased in a broad region of pH2O = 2~43 kPa, whereas its photocatalytic system and also opens up an intriguing way to enhance
dependence on the partial pressure of CO2 was only observed in a the efficiency and selectivity for CO2 conversion.
limited region (≤ 2 kPa). The results reflected that higher H2O coverage
favored its activation on the catalyst surface, which played a dominant CRediT authorship contribution statement
role in the reaction with the adsorbed CO2 molecules, as illustrated in
Fig. 1. Thus the kinetically sluggish water oxidation reaction could be Xiantao Hu: Experiment on photocatalysis, Writing - Original Draft.
dramatically promoted. As further evidence, Fig. 7c presented a much Zhanjun Xie: Experiment on photothermal effect, Writing - Original
higher O2 production rate (41.3 μmol g− 1 h− 1) in TiO2+PM system. The Draft. Qian Tang: Data analysis, Writing - Original Draft. Heng Wang:
O2 production occurred in a 4-electron transfer process, so the electrons Investigation. Lianbin Zhang: Helpful discussion, Writing - Review &
from water oxidation reaction are equivalent to the those that were Editing. Jingyu Wang: Conceptualization, Writing - Review & Editing,
consumed during CO2 reduction. The simultaneous monitoring of CH4, Supervision.
CO and O2 production verifies the overall redox reaction over TiO2
photocatalyst with assitence of PM-induced water steam generation, Declaration of Competing Interest
achieving enhanced effiicency for CO2 reduction and H2O oxidation,
much higher than the reported values over various kinds of TiO2 and The authors declare that they have no known competing financial
other photocatalysts in literatures (Table S2). The theoretical study interests or personal relationships that could have appeared to influence
discovered that the protons were transferred concertedly with two the work reported in this paper.
electrons to the linear CO2 to form a CO molecule, which was strongly
adsorbed and further reduced to CH4 on anatase TiO2 surface [62]. The Acknowledgements
CH4 production could be promoted by the additional introduction CO
(Fig. S15). As a result, the mechanism for CO2 reduction involved We thank the Analysis and Testing Centre, Huazhong University of
multiple steps, (1) the adsorbed H2O being dissociated to produce pro­ Science and Technology for the characterization of materials. This study
tons as the first step, (2) the proton being transferred to CO2 and the was supported by the National Natural Science Foundation of China
protonated CO2 being decomposed to CO, (3) the strongly adsorbed CO (21771070, 22122602) and the Fundamental Research Funds for the
being hydrogenated and deoxygenation to form CH4 as final product. Central Universities (2018KFYYXJJ120, 2019KFYRCPY104).

4. Conclusion
Appendix A. Supplementary data

In summary, the present work focused on the impact of H2O vapor on


Supplementary material related to this article can be found, in the
the photocatalytic CO2 reduction. It has been demonstrated that the CO2
online version, at doi:https://doi.org/10.1016/j.apcatb.2021.120635.
conversion efficiency strongly depends on the partial pressure of H2O
vapor, whereas its dependence on the partial pressure of CO2 was only
References
observed in a limited region (≤ 2 kPa). Thanks to a photothermal effect
of PPy membrane, the well-designed TiO2+PM system accelerated H2O [1] X. Li, J. Yu, M. Jaroniec, X. Chen, Chem. Rev. 119 (2019) 3962–4179.
evaporation to increase the reaction temperature at surface catalytic [2] T. Kong, Y. Jiang, Y. Xiong, Chem. Soc. Rev. 49 (2020) 6579–6591.
sites from 73 to 86 ◦ C and partial pressure of H2O vapor from 17 to 43 [3] Z.-W. Wang, Q. Wan, Y.-Z. Shi, H. Wang, Y.-Y. Kang, S.-Y. Zhu, S. Lin, L. Wu, Appl.
Catal. B-Environ. 288 (2021), 120000.

7
X. Hu et al. Applied Catalysis B: Environmental 298 (2021) 120635

[4] D. Li, A. Hussain, Y. Wang, C. Huang, P. Li, M. Wang, T. He, Appl. Catal. B-Environ. [33] H. Wu, X. Kong, X. Wen, S.-P. Chai, E.C. Lovell, J. Tang, Y.H. Ng, Angew. Chem.
286 (2021), 119887. Int. Ed. 60 (2020) 8455–8459.
[5] W. Zhang, A.R. Mohamed, W.-J. Ong, Angew. Chem. Int. Ed. 59 (2020) [34] Q. Yang, M. Luo, K. Liu, H. Cao, H. Yang, Appl. Catal. B-Environ. 276 (2020),
22894–22915. 119174.
[6] H. Wang, Q. Tang, Z. Chen, T. Li, J. Wang, Sci. China-Mater. 63 (2020) 2189–2205. [35] Y. Ma, Z. Wang, X. Xu, J. Wang, Chin. J. Catal. 38 (2017) 1956–1969.
[7] B. Ma, G. Chen, C. Fave, L. Chen, R. Kuriki, K. Maeda, O. Ishitani, T.-C. Lau, [36] X. Jiao, X. Li, X. Jin, Y. Sun, J. Xu, L. Liang, H. Ju, J. Zhu, Y. Pan, W. Yan, et al.,
J. Bonin, M. Robert, J. Am. Chem. Soc. 142 (2020) 6188–6195. J. Am. Chem. Soc. 139 (2017) 18044–18051.
[8] T. Nakajima, Y. Tamaki, K. Ueno, E. Kato, T. Nishikawa, K. Ohkubo, Y. Yamazaki, [37] A. Sarkar, E. Gracia-Espino, T. Wagberg, A. Shchukarev, M. Mohl, A.-R. Rautio,
T. Morimoto, O. Ishitani, J. Am. Chem. Soc. 138 (2016) 13818–13821. O. Pitkanen, T. Sharifi, K. Kordas, J.-P. Mikkola, Nano Res. 9 (2016) 1956–1968.
[9] M.F. Kuehnel, K.L. Orchard, K.E. Dalle, E. Reisner, J. Am. Chem. Soc. 139 (2017) [38] K. Mori, H. Yamashita, M. Anpo, RSC Adv. 2 (2012) 3165–3172.
7217–7223. [39] L. Lu, B. Wang, S. Wang, Z. Shi, S. Yan, Z. Zou, Adv. Funct. Mater. 27 (2017),
[10] B. Han, X. Ou, Z. Deng, Y. Song, C. Tian, H. Deng, Y.-J. Xu, Z. Lin, Angew. Chem. 1702447.
Int. Ed. 57 (2018) 16811–16815. [40] Y. Ma, Q. Tang, W.-Y. Sun, Z.-Y. Yao, W. Zhu, T. Li, J. Wang, Appl. Catal. B-
[11] D. Saito, Y. Yamazaki, Y. Tamaki, O. Ishitani, J. Am. Chem. Soc. 142 (2020) Environ. 270 (2020), 118856.
19249–19258. [41] F. Zhao, X. Zhou, Y. Shi, X. Qian, M. Alexander, X. Zhao, S. Mendez, R. Yang, L. Qu,
[12] H. Zhang, Y. Wang, S. Zuo, W. Zhou, J. Zhang, X.W.D. Lou, J. Am. Chem. Soc. 143 G. Yu, Nat. Nanotechnol. 13 (2018) 489–495.
(2021) 2173–2177. [42] C. Chen, Y. Li, J. Song, Z. Yang, Y. Kuang, E. Hitz, C. Jia, A. Gong, F. Jiang, J.
[13] P.V. Kamat, S. Jin, ACS Energy Lett. 3 (2018) 622–623. Y. Zhu, et al., Adv. Mater. 29 (2017), 1701756.
[14] H. Pang, X. Meng, P. Li, K. Chang, W. Zhou, X. Wang, X. Zhang, W. Jevasuwan, [43] H. Liang, Q. Liao, N. Chen, Y. Liang, G. Lv, P. Zhang, B. Lu, L. Qu, Angew. Chem.
N. Fukata, D. Wang, et al., ACS Energy Lett. 4 (2019) 1387–1393. Int. Ed. 58 (2019) 19041–19046.
[15] M. Ghoussoub, M. Xia, P.N. Duchesne, D. Segal, G. Ozin, Energy Environ. Sci. 12 [44] L. Zhu, T. Ding, M. Gao, C.K.N. Peh, G.W. Ho, Adv. Energy Mater. 9 (2019),
(2019) 1122–1142. 1900250.
[16] M. Schreck, M. Niederberger, Chem. Mater. 31 (2019) 597–618. [45] L. Sun, Z. Li, R. Su, Y. Wang, Z. Li, B. Du, Y. Sun, P. Guan, F. Besenbacher, M. Yu,
[17] Y. Dou, A. Zhou, Y. Yao, S.Y. Lim, J.-R. Li, W. Zhang, Appl. Catal. B-Environ. 286 Angew. Chem. Int. Ed. 57 (2018) 10666–10671.
(2021), 119876. [46] G. Chen, J. Sun, Q. Peng, Q. Sun, G. Wang, Y. Cai, X. Gu, Z. Shuai, B.Z. Tang, Adv.
[18] Y. Liu, D. Shen, Q. Zhang, Y. Lin, F. Peng, Appl. Catal. B-Environ. 283 (2021), Mater. 32 (2020), 1908537.
119630. [47] M. Xu, X. Hu, S. Wang, J. Yu, D. Zhu, J. Wang, J. Catal. 377 (2019) 652–661.
[19] C. Xia, H. Wang, J.K. Kim, J. Wang, Adv. Funct. Mater. 31 (2020), 2008247. [48] T. Kajiwara, M. Fujii, M. Tsujimoto, K. Kobayashi, M. Higuchi, K. Tanaka,
[20] J. Low, B. Dai, T. Tong, C. Jiang, J. Yu, Adv. Mater. 31 (2019), 1802981. S. Kitagawa, Angew. Chem. Int. Ed. 55 (2016) 2697–2700.
[21] P. Madhusudan, R. Shi, S. Xiang, M. Jin, B.N. Chandrashekar, J. Wang, W. Wang, [49] H. Zhong, Z. Hong, C. Yang, L. Li, Y. Xu, X. Wang, R. Wang, ChemSusChem 12
O. Peng, A. Amini, C. Cheng, Appl. Catal. B-Environ. 282 (2021), 119600. (2019) 4493–4499.
[22] Y. Huang, K. Wang, T. Guo, J. Li, X. Wu, G. Zhang, Appl. Catal. B-Environ. 277 [50] H. Rao, C.-H. Lim, J. Bonin, G.M. Miyake, M. Robert, J. Am. Chem. Soc. 140 (2018)
(2020), 119232. 17830–17834.
[23] H. Wang, X. Hu, Y. Ma, D. Zhu, T. Li, J. Wang, Chin. J. Catal. 41 (2020) 95–102. [51] S. Xie, Y. Wang, Q. Zhang, W. Deng, Y. Wang, ACS Catal. 4 (2014) 3644–3653.
[24] J. Wang, M. Xu, J. Zhao, H. Fang, Q. Huang, W. Xiao, T. Li, D. Wang, Appl. Catal. B- [52] Z. Fu, X. Wang, A.M. Gardner, X. Wang, S.Y. Chong, G. Neri, A.J. Cowan, L. Liu,
Environ. 237 (2018) 228–236. X. Li, A. Vogel, et al., Chem. Sci. 11 (2020) 543–550.
[25] S. Huang, T. Ouyang, B.-F. Zheng, M. Dan, Z.-Q. Liu, Angew. Chem. Int. Ed. 60 [53] M. Tahir, N.S. Amin, Appl. Catal. A-Gen. 467 (2013) 483–496.
(2021) 9546–9552. [54] M. Tahir, N.S. Amin, Chem. Eng. J. 230 (2013) 314–327.
[26] S. Huang, B.-F. Zheng, Z.-Y. Tang, X.-Q. Mai, T. Ouyang, Z.-Q. Liu, Chem. Eng. J. [55] D.C. Sorescu, J. Lee, W.A. Al-Saidi, K.D. Jordan, J. Chem. Phys. 137 (2012),
422 (2021), 130086. 074704.
[27] S. Wang, M. Xu, T. Peng, C. Zhang, T. Li, I. Hussain, J. Wang, B. Tan, Nat. Commun. [56] X. Chang, T. Wang, J. Gong, Energy Environ. Sci. 9 (2016) 2177–2196.
10 (2019) 676. [57] G.E.M. Schukraft, R.T. Woodward, S. Kumar, M. Sachs, S. Eslava, C. Petit,
[28] X. Chang, T. Wang, J. Gong, Energy Environ. Sci. 9 (2016) 2177–2196. ChemSusChem 14 (2021) 1720–1727.
[29] Q. Tang, Y. Ma, J. Wang, Sol. RRL 5 (2021), 2000443. [58] Z. Han, Y. Fu, Y. Zhang, X. Zhang, X. Meng, Z. Zhou, Z. Su, Dalton Trans. 50 (2021)
[30] H. Dong, X. Zhang, Y. Lu, Y. Yang, Y.-P. Zhang, H.-L. Tang, F.-M. Zhang, Z.- 3186–3192.
D. Yang, X. Sun, Y. Feng, Appl. Catal. B-Environ. 276 (2020), 119173. [59] H. Li, C. Rameshan, A.V. Bukhtiyarov, I.P. Prosvirin, V.I. Bukhtiyarov,
[31] X. Yu, Z. Yang, B. Qiu, S. Guo, P. Yang, B. Yu, H. Zhang, Y. Zhao, X. Yang, B. Han, G. Rupprechter, Surf. Sci. 679 (2019) 139–146.
et al., Angew. Chem. Int. Ed. 58 (2019) 632–636. [60] M.T. Nayakasinghe, Y. Han, N. Sivapragasam, D.S. Kilin, U. Burghaus, Chem.
[32] G. Singh, J. Lee, A. Karakoti, R. Bahadur, J. Yi, D. Zhao, K. AlBahily, A. Vinu, Commun. 54 (2018) 9949–9952.
Chem. Soc. Rev. 49 (2020) 4360–4404. [61] Z. Gao, Y. Ding, J. Mol. Model. 23 (2017) 187.
[62] Y. Ji, Y. Luo, ACS Catal. 6 (2016) 2018–2025.

You might also like