Open access

Photocatalytic Hydrogenation on Semiconductor Particles

Written By

Shigeru Kohtani, Eito Yoshioka and Hideto Miyabe

Submitted: 19 December 2011 Published: 10 October 2012

DOI: 10.5772/45732

From the Edited Volume

Hydrogenation

Edited by Iyad Karamé

Chapter metrics overview

6,134 Chapter Downloads

View Full Metrics

1. Introduction

Photocatalytic hydrogenation on semiconductor particles is a quite unique methodology compared to the conventional hydrogenations such as catalytic hydrogenation on metals or homogeneous catalysis using metal complexes. The catalysis can be defined as a light-driven redox reaction at a solid/liquid or a solid/gas interface. The semiconductor photocatalysis have been mainly developed by researchers in the areas of photochemistry, electrochemistry, and heterogeneous catalysis. Since Fujishima and Honda have discovered the photoelectrochemical water splitting upon titanium dioxide (TiO2) photoelectrode in the early 1970’s [1], the heterogeneous photocatalysis, especially using TiO2, has been applied to various fields such as storage of solar energy [2,3], environmental purification [3,4], and specific surface properties of self-cleaning, self-sterilizing, and anti-fogging induced by incident light [5]. The photocatalysis has also attracted much attention with respect to inducing characteristic organic transformations since 1970’s [6-11]. Among those, the photocatalytic hydrogenation of ethene and ethyne on TiO2 was already reported in 1975 [12]. Since then, the reductive photocatalysis has been applied to the hydrogenation for several organic compounds having various double or triple bonds.

Herein, we summarize the progress in photocatalytic hydrogenations covering the literatures available up to 2011. At first, we introduce the fundamentals of semiconductor photocatalysis in order to understand the mechanistic principles. Next, we review the reports on applications to the photocatalytic hydrogenation. Here, we refer to scope and limitation of the photocatalytic hydrogenation on semiconductor particles. Advantages and disadvantages using this method are also compared to those of the conventional hydrogenations. Finally, prospects of the photocatalytic hydrogenation are discussed.

Advertisement

2. Fundamentals

Some metal oxides (e.g. titanium dioxide (TiO2) etc.) and sulfides (e.g. cadmium sulfide (CdS), zinc sulfide (ZnS) etc.) are regarded as semiconductors, in which electrons (e-) photogenerated in conduction band (CB) and holes (h+) simultaneously generated in valence band (VB) play important roles in electroconductivity as well as chemical reactivities on the surface, i.e. e- and h+ can induce redox reactions as shown in Fig. 1. This reaction system is so called “a micro-photoelectrochemical cell”, consisting of an anode and a cathode within a small particle. The hydrogenation generally proceeds as the reduction of an electron acceptor (A) followed by the protonation as depicted in Fig. 1. The photocatalytic reduction of A can be carried out in the presence of a large excess amount of electron donor (D) such as alcohols or amines, and in the absence of molecular oxygen (O2). The aim using the electron donor (D) is to scavenge hole generated in VB, thereby diminishing the degree of recombination between e- and h+ within the semiconductor particles. Thus, it is important for the reductive hydrogenation of organic substrates to choice an appropriate electron donor (D) as hole scavenger. Moreover, it is necessary to purge O2 gas from the reaction system in order to improve the reduction efficiency, since O2 acts as a competitive electron acceptor.

Figure 1.

Mechanistic principle of photocatalytic hydrogenation on a semiconductor particle

Important points in the semiconductor photocatalyst materials are band gap energy between CB and VB and potential energy levels of CB and VB. The band gap energy defines the minimum photon energy absorbed by the semiconductor materials: Band gap (eV) = 1240/wavelength (nm). Therefore, if a photocatalyst possesses less than 3.1 eV band gap, it can absorb photons with visible light region (> 400 nm). Fig. 2 indicates the band gaps and the band levels (V vs. standard hydrogen electrode (SHE) at pH 7) of TiO2 (anatase), TiO2 (rutile), ZnS, and CdS semiconductor materials as measured photoelectrochemically [8]. TiO2 has four polymorphs: rutile (tetragonal, the most stable phase), anatase (tetragonal), brookite (orthorhombic), and TiO2 (B) (monoclinic), in which rutile and anatase have been mainly used as photocatalysts. The band gap of anatase (3.2 eV) is slightly larger than that of rutile (3.0 eV), because the CB level of anatase is located at 0.2 eV more negative than that of rutile as depicted in Fig. 2. In order to proceed the photocatalytic reaction effectively, the bottom level of CB has to be more negative than a reduction potential of A (A/A-・), while the top level of VB has to be more positive than an oxidation potential of D (D+・/D) as shown in Fig. 2.

Figure 2.

Band gaps and band edge positions of some semiconductor photocatalysts

It should be also noted that the band levels usually shift with a change in pH (-0.059 V/pH) for oxide materials. In addition to this pH dependence, surface impurities, adsorbed compounds, and the change to organic solvents would induce band shifts. For an example, the negative band shift is observed when water is replaced by organic solvents such as acetonitrile and alcohols [13]. The negative CB level is essential for the photoreductive hydrogenation reaction. From this point of view, ZnS and CdS having the negative CB level are effective candidates for the photocatalytic hydrogenation. However, these sulfides can be oxidized by h+ in the absence of an appropriate hole scavenger: CdS or ZnS + 2h+ → Cd2+ or Zn2+ + S, which is so called “photocorrosion” [8]. Thus, the choice of a suitable hole scavenger (solvent) is particularly important for hydrogenation using ZnS or CdS.

It is important to estimate a quantum yield (QY) in photocatalytic reactions. QY is defined as the number of events which occurs per photon absorbed by a photocatalyst as follows:

QY=(number of reacted electrons or holes)(number of absorbed photons)E1

However, it is difficult to determine the real number of photons absorbed by a photocatalyst in a dispersed system because of light scattering by the photocatalyst powder. Therefore, in general, an apparent QY as described below is applied to the actual photocatalytic reactions.

Apparent  QY=(number of reacted electrons or holes)(number of incident photons)E2

The apparent QY is estimated to be smaller than the real QY because the number of absorbed photons is usually smaller than that of incident photons.

Loading of fine particles, usually nanometer size of noble metals (e.g. platinum (Pt) or silver (Ag) etc.) or metal oxide (nickel oxide (NiO) etc.), on the photocatalyst surface often improves the reaction efficiency as a co-catalyst due to the following reasons: (1) the particles enhance the charge separation between e- and h+ and prevent the charge recombination, and (2) the particles act as effective reactive sites in the photocatalytic transformation processes. For an example, Pt fine particles loaded on photocatalyst greatly improve the chemical efficiency of hydrogen evolution from water or alcohol under irradiation.

Advertisement

3. Applications

3.1. Hydrogenation of alkenes and alkynes

Photocatalytic hydrogenation of ethene and ethyne firstly reported by Boonstra and Mutsaers in 1975 [12]. Some hydrogenated products have been found upon the UV illumination (320 – 390 nm) of TiO2 in an atmosphere of ethene or ethyne. Main products from ethene were methane, ethane, propane, and n-butane, whereas, in the case of ethyne, the products were methane, ethene, ethane, and propane. In these reactions, the hydrogen source has been thought to be the surface Ti-OH group on TiO2.

Anpo and co-workers investigated the photocatalytic hydrogenation of various alkenes and alkynes in the presence of water vapor upon TiO2 [14-17]. Major products on the UV illuminated TiO2 powders were the compounds formed by hydrogenation accompanied by C=C or C≡C bond fission [14, 15]. In contrast, a significant enhancement of hydrogenation products without the bond fission has been observed on the UV illuminated Pt-loaded TiO2 (Pt/TiO2) powders [16]. Especially, the Pt-loaded rutile TiO2 predominantly catalyzed the hydrogenation of propyne (CH3C≡CH) to afford C3H6 and C3H8 (C3/C2 ratio = 5.60) as listed in Table 1. Anpo et al. further investigated the size effects on Pt/TiO2 particles in the photocatalytic hydrogenation of CH3C≡CH with H2O [17]. Quantum yields of the whole photoreactions became smaller with increasing the particle size of Pt/TiO2 (rutile), whereas selectivity of the hydrogenation products without vs. with the bond fission (C3H8/C2H6) reversely became larger and reached at C3H8/C2H6 = 99 upon the particle size of 200 nm.

Catalysts TypeProducts (10-9 mol m-2 h-1)C3/C2 ratio
CH4C2H6C2H4C3H6C3H8
TiO2 (P25a)1.9012.30.401.100.500.12
Ptb/TiO2 (P25a)2.4015.70.5015.70.400.86
TiO2 (rutile)0.553.270.100.480.190.17
Ptb/TiO2 (rutile)0.221.600.2310.31.105.60

Table 1.

The amount of products formed in the photocatalytic hydrogenation of CH3C≡CH on the four type of TiO2 in the presence of water vapor at 300 K [16].

The hydrogenation of C-C multiple bonds of alkenes and alkynes was also examined in alcohols using Pt/TiO2 powders by Yamataka et al. [18]. In this case, some saturated alkanes were produced in good yields (> 50% after 24 h), while the alcoholic solvents were concurrently oxidized to the corresponding carbonyl compounds. Baba et al. have synthesized bimetal-loaded TiO2 such as Pd/TiO2/Ni or Pt/TiO2/Cu and applied those to the hydrogenation of ethene [19]. The hydrogenation to ethane was efficiently occurred upon the bimetal/TiO2, on which the role of the latter metals (Ni or Cu) was to suppress the hydrogen production as side reaction. Titanium-silicon (Ti/Si) binary oxides were prepared and utilized for the photocatalytic hydrogenation of alkenes and alkynes with H2O by Anpo’s group [20-22]. The reactivity and selectivity for the hydrogenation of CH3C≡CH were investigated as a function of the Ti content [21, 22]. It was found that the hydrogenation with bond fission producing C2H6 and CH4 was predominant in regions of low Ti content, whereas the hydrogenation yielding C3H6 proceeded in regions of high Ti content. They have revealed that tetrahedrally coordinated titanium oxide species played a significant role in the efficient photoreaction with a high selectivity for the hydrogenation with the bond fission, while the catalysts involving the aggregated and octahedrally coordinated titanium dioxide species showed a high selectivity for the hydrogenation producing C3H6, being similar to reaction using the powdered TiO2 catalysts. Molybdenum oxide or sulfide complexes supported on TiO2 were developed by Kunts and applied to the photocatalytic hydrogenation of ethyne [23, 24]. The sulfur systems were somewhat more efficient in the hydrogenation, and also favored the 2 electron-transferred product ethene rather than the 4 electron-transferred product ethane.

Yanagida et al. reported that visible light response CdS nanocrystallites (2 – 5 nm) catalyzed the efficient hydrogenation of electron-deficient alkenes with triethylamine (TEA) (Table 2) [25]. This reaction is the first example of visible light induced hydrogenation of alkenes on semiconductor particles. The hydrogenation (two electron reduction) was always accompanied with cis-trans isomerization (one electron reduction /oxidation process). The yield of the hydrogenation becomes favorable with increasing light intensity [26]. TEA was used for a sacrificial electron donor, in which the oxidation reaction proceeded as follows: Et3N + H2O + 2h+ → Et2NH + CH3CHO + 2H+.

R1R2Time / hConversion / %Yield / %-Ereda / V
CO2CH3CO2CH3 (cis)2100701.56
CO2CH3CO2CH3 (trans)2100601.60
p-CNC6H4CN (cis)370471.75
p-CNC6H4CN (trans)368411.73
p-CNC6H4CO2CH3 (cis)3100761.75
p-CNC6H4CO2CH3 (trans)3100921.75
C6H5CO2CH3 (trans)310trace1.98

Table 2.

CdS-catalyzed photohydrogenation of alkenes with triethylamine in MeOH [25].

3.2. Hydrogenation of imine, azo, and azide compounds

Hydrogenation of the C=N bond of imine intermediates upon the Pt/TiO2 and CdS photocatalysts was observed during some inter- or intramolecular deaminocondensations in one-pot reaction systems as reported by Ohtani and co-workers [27, 28]. Fig. 3 illustrates the photocatalytic transformation of primary amine to symmetrical or asymmetrical secondary amines via the photocatalytic deaminocondensation in water or alcohols. In water [27], a primary amine was firstly oxidized by two holes to form an imine (R1CH=NH) followed by hydrolysis to give a corresponding aldehyde (R1CHO). Next, the aldehyde was condensed with another amine to yield the imine intermediate (R1CH=NCH2R1). Finally, the imine was photocatalytically hydrogenated to produce the symmetrical secondary amine (R1CH2NHCH2R1). On the other hand, in alcoholic solvent [28], the alcohol molecule (sacrificial hole scavenger) was primarily oxidized by two holes to afford the corresponding carbonyl compound (R2R3C=O) which was also condensed with amine to yield the imine intermediate (R2R3C=NCH2R1). The imine was further hydrogenated to form the asymmetrical secondary amine (R2R3CHNHCH2R1). This mechanistic principle has been used for several inter- or intramolecular reaction systems as depicted in Scheme 1. Furthermore, photocatalytic deaminocyclization of L-lysine to pipecolinic acid was examined using TiO2 and CdS, and concluded that these two photocatalysts exhibited the different stereochemistry [29, 30].

Kisch el al. found that photocatalytic hydrogenation of azobenzene to hydrazobenzen on ZnS or CdS proceeded as a side reaction of a photocatalyzed addition of 3,4-dihydropyran to azobenzene giving 1-(3,4-Dihydro-2H-pyran-4-yl)-1,2-diphenylhydrazine (DPDH) (Scheme 2) [8, 31]. The formation of the hydrazobenzen was strongly favored when Pt-loaded ZnS or Pt-loaded CdS was used as the photocatalyst. This would be caused by the two electron-transfer process preferentially occurred on the platinum fine particles.

Figure 3.

Deaminocondensation reactions in one pot system containing hydrogenation process.

Scheme 1.

Inter- (A) or intramolecular (B) deaminocondensation reactions [27].

Scheme 2.

Pt-loading effect on photocatalytic hydrogenation and addition reactions upon ZnS or CdS photocatalysts [8, 31].

Photocatalytic hydrogenation of aromatic azides (Ar-N3) to amines upon CdS and cadmium selenide (CdSe) was examined by Warrier et al. [32]. The wide scope of the reaction was confirmed with compounds containing electron-withdrawing (-NO2, COOR, COR) and electron-donating groups (-OMe, -R, -Cl) at the para-, meta-, and ortho-positions. These reactions took place with high quantum yields (near 0.5). Sodium formate was used as electron donor which was oxidized to CO2 during the reaction.

3.3. Hydrogenation of carbonyl compounds

Aldehydes and ketones are two classes of compounds that possess a reactive carbonyl group. The first report on photocatalytic hydrogenation of carbonyl compound is, to our knowledge, the hydrogenation of pyruvate to lactate under irradiation of aqueous suspension of TiO2 reported by Cuendet and Grätzel in 1987 [33]. Later, Li and co-workers reported P25 TiO2-catalyzed hydrogenation of benzaldehyde forming benzylalcohol in ca. 80% yield [34]. This reaction was recently applied to a micro-reaction system by Matsushita et al. [35, 36]. Kohtani et al. recently demonstrated that the P25 TiO2 powder exhibited the excellent photocatalytic activity to hydrogenate several aromatic ketones into corresponding secondary alcohols under the combination of UV light irradiation and deaerated conditions in ethanol as summarized in Table 3 [37]. Acetaldehyde was simultaneously produced in the oxidation of ethanol by h+ generated in the VB or those trapped at the surface sites on TiO2. The desired secondary alcohols were obtained in almost quantitative yields for more than ten examples: e.g. acetophenone (Ar = C6H5, R = Me) to 1-phenylethanol in 97% yield (highlighted in gray color in Table 3). They also found that most of the reaction rates depend on the reduction potential (Ered) of substrates, except for 2,2,2-trifluoroacetophenone [37]. In general, aldehydes are more reactive than ketones because of electronic and steric factors. Electrochemically, Ered of aldehydes are more positive than those of ketones. However, the hydrogenations of aldehydes were accompanied with competitive formation of by-products such as diethyl acetal etc., leading to erosion of chemical efficiency.

Aliphatic aldehydes and ketones were hydrogenated by using the ZnS photocatalyst, because the CB level of ZnS is sufficiently negative as indicated in Fig. 2 [38, 39]. In contrast, the photocatalytic hydrogenation did not occur upon TiO2, since the CB level of TiO2 is too positive to reduce the aliphatic ketones [37]. Yanagida and co-workers reported that acetaldehyde was reduced and oxidized upon the UV irradiated ZnS nano-crystallites (2 – 5 nm) in water as depicted in Scheme 3 [38]. The reduction products were ethanol and H2 generated by the photocatalytic reduction of H2O, while acetaldehyde was oxidized by the photogenerated holes to produce acetic acid, diacetyl, and acetoin. The apparent quantum yield of ethanol (main product) formation was 0.25 at 313 nm. The similar result was observed in the photocatalytic reaction of propionaldehyde on the ZnS nano-crystallite [38]. The ZnS nano-crystallites were further applied to the photocatalytic hydrogenation of aliphatic ketones such as acetone, 2-butanone, 3-pentanone, 2-hexanone, cyclopentanone, and cyclohexanone in the presence of sacrificial hole scavengers of S2- and SO32- (Oxidation reaction: S2- + SO32- +2h+ → S2O32-) [39]. Most of examined ketones were photocatalytically hydrogenated at comparable rates to give the alcohols in almost quantitative yields except for acetone. The apparent quantum yield for the formation of 2-butanol was 0.27 at 313 nm.

Yanagida et al. applied the visible light response CdS nanocrystallite to the photoreduction of aromatic ketones [25, 26, 40]. Two types of the reduction products, secondary alcohols from two electron-transfer process and pinacols from one electron-transfer process, were observed as listed in Table 4, while the oxidation of TEA as a sacrificial electron donor by the VB holes afforded diethylamine and acetaldehyde. The yields of the hydrogenation involving two electron-transfer process were preferable for substrates possessing the electron- withdrawing group (-CN or –Cl) as shown in Table 4, and further became favorable with increasing light intensity [26] and with decreasing the particle size [40].

Scheme 3.

Photocatalytic reduction and oxidation reactions of acetaldehyde in water on ZnS.

Substrates (1)Time (h)Conversion
(%)
Yield of 2 (%)Yield of 3 (%)-Ered (V)b
ArR
C6H5H110076±2571.99
C6H5Me410097±2"/>992.13
C6H5Et89187"/>992.18
C6H5i-Pr82425322.21
C6H5t-Bu8177422.24
C6H5CF387063101.59
C6H5C6H548278971.83
2-MeC6H4Me86663842.16
4-MeC6H4Me8100"/>99"/>992.15
2-FC6H4Me310099"/>991.86
3-FC6H4Me310097"/>992.04
4-FC6H4Me4100"/>99"/>992.15
2,6-F2C6H4Me51009694-
C6F5Me2100"/>99"/>991.83
1-C10H7H19576±3771.66
1-C10H7Me10100"/>99931.86
2-C10H7H19432±7931.70
2-C10H7Me510094±4"/>991.73
8100"/>99"/>992.06
810091"/>99-
28281981.42

Table 3.

Photocatalytic hydrogenation of aromatic carbonyl compounds on P25 TiO2 [37].a

Substrates (4)Time (h)Conversion of 4 (%)Yield of 5 (%)Yield of 6 (%)-Ered (V)a
R1R2
p-CNC6H4C6H521009501.17
p-ClC6H4C6H539870121.32
p-ClC6H4p-ClC6H43989041.35
C6H5C6H531009551.55
p-CH3OC6H4C6H539733561.56
p-CH3OC6H4p-CH3OC6H435215401.56
C6H5CH3310trace802.00

Table 4.

Photocatalytic hydrogenation of aromatic ketones on nanocrystallized CdS [25].

1,2-Diketones such as camphorquinone, 1-phenyl-1,2-propanedione, and benzil were hydrogenated to the corresponding α-hydroxyketones in moderate to good yields on the UV-irradiated P25 TiO2 as shown in Scheme 4 [41, 42]. The yields and stereoselectivities were increased in the presence of water or TEA as a sacrificial electron donor in methanol solvent. The endo-hydroxycamphors were formed much more favorably than the exo-products, though there was little selectivity between 2 and 3 positions.

Scheme 4.

Photocatalytic hydrogenations of diketone compounds upon the P25 TiO2 powder.

3.4. Hydrogenation of carbon dioxide

With increasing concerns about rising atmospheric CO2 concentration, the need for research to utilize CO2 in chemical synthesis has been increased greatly [43]. Therefore, the photocatalytic hydrogenation of CO2 upon semiconductor powders has received much attention [9-11, 44-56]. Among the earliest studies on the photocatalytic hydrogenation of CO2, Inoue et al. examined a wide range of semiconductors (WO3, TiO2, ZnO, CdS, GaP, and SiC in 200 – 400 mesh) in aqueous solution [44]. The results indicated that CO2 was reduced to HCOOH, HCHO, and CH3OH, and the product yields correlated to the position of the CB level of the semiconductors. In the case of WO3, since the CB level was insufficiently negative to reduce CO2, no reduction products were observed. In contrast, SiC having the most negative CB level gave the highest yields. Later, Henglein and Gutiérrez reported that high quantum yield of 0.20 was achieved in the hydrogenation of CO2 to formate using ZnS colloidal suspension under UV light irradiation [45]. Here, the solvent was a mixture of water and ethanol, or 2-propanol as the hole scavenger. After the earliest works, a great deal of effort has been devoted to studying the photocatalytic reduction of CO2. Most of works are summarized in recent reviews [9-11], in which the use of TiO2, SrTiO3, ZrO, ZnS, or CdS is reported. Recently, visible-light-driven new materials were developed and applied to the photocatalytic reduction of CO2 under visible light irradiation: for examples, InTaO4 [46], LaCoO3 [47], BiVO4 [48], ZnGaO4 [49], Zn2GeO4 [50, 51], Bi2WO6 [52] etc.

The photocatalytic reduction of CO2 on semiconductor powders gives several reduction products such as CO, HCOOH, HCHO, CH3OH, CH4, and various hydrocarbons. Among those, methanol is the most valuable product because it can be directly used as a fuel or a building block. High efficiency and selectivity in the preparation of methanol were obtained by the use of Ti-oxide/Y-zeolite catalysts containing highly dispersed tetrahedral titanium oxide species under UV irradiation [53, 54]. The charge-transfer excited state of these species played an important role in the selectivity for producing CH3OH, in contrast to the different selectivity giving CH4 on bulk TiO2 photocatalyst. Ti-incorporated mesoporous silica also exhibited high activity in the photoreduction of CO2 with water to generate CH3OH and CH4 under UV irradiation [55]. Methanol was also selectively produced by employing CaFe2O4 [56] and NiO-loaded InTaO4 [47] under visible light irradiation.

3.5. Hydrogenation of Nitrate and nitrite ions, and nitroaromatics

The use of artificial fertilizers in agriculture has caused a great deal of concern regarding water pollution induced by production of nitrate (NO3-) and nitrite (NO2-) ions from the fertilizers [57]. Therefore, chemical processes for elimination or hydrogenation of NO3- ions have been extensively studied photocatalytically. TiO2 photocatalyst has been used for the hydrogenation of NO3- and NO2- ions to ammonia [58-60]. Other photocatalysts such as SrTiO3 [61], K4NbO17 [61], tantalate oxides (K3Ta3Si2O13, BaTa2O6, KTaO3, NaTaO3) [62], and ZnS [63], have also been reported as active photocatalysts for the hydrogenation of nitrate ions under UV light irradiation. It is worth noting that Kudo et al. developed a 0.1% nickel doped ZnS (Zn0.999Ni0.001S) photocatalyst to extend the photocatalytic response of ZnS toward visible region [64], and applied it into the reduction of NO3- and NO2- ions under visible light irradiation [65]. The amounts of products and electrons consumed in the reduction of NO3- and NO2- are listed in Table 5. The ratios of the amounts of NH3/NO2- and NH3/N2 increased in the presence of the platinum cocatalyst, indicating that selectivity for the NH3 production was improved by the Pt loading on Zn0.999Ni0.001S. More recently, Yamauchi and co-workers reported highly selective ammonia synthesis from nitrate with photocatalytically generated hydrogen on CuPd nanoalloys loaded on TiO2, on which ammonia was selectively produced up to 78% yield with hydrogen evolution under UV light irradiation [66].

Catalysts TypeReactantAmounts of products (electron consumed)/μmol
(initial conc./ molL-1)H2 bNO2-NH3N2
non-loadedno NO3- nor NO2- (0)214 (428)(-)(-)(-)
non-loadedNO3- (1.0)60 (120)250 (500)c21 (168)d2.7 (27)e
1wt% Pt-loadedNO3- (1.0)212 (424)14 (28)c11 (88)d0.79 (7.9)e
non- loadedNO2- (0.01)37 (74)(-)41 (246)f2.4 (14)g
1wt% Pt-loadedNO2- (0.01)96 (192)(-)65 (390)f9.5 (57)g

Table 5.

Reduction of NO3- and NO2- on Zn0.999Ni0.001S photocatalysts [65].a

Several organic nitroaromatics can be easily hydrogenated to afford corresponding amino compounds in the presence of sacrificial hole scavenger upon the UV irradiated TiO2 as firstly reported by Li and co-workers in 1993 [67]. Since then, the photocatalytic hydrogenation of nitro compounds using some kinds of semiconductor photocatalysts was extensively studied by a number of researchers. The hydrogenation reactions proceeded almost quantitatively. The details are summarized in recent reviews [9-11]. Recent progress in this research has been directed toward the development of visible light response photocatalysts such as nitrogen-doped TiO2 [68] and dye-sensitized TiO2 loaded with transition metal nanoparticles [69].

Advertisement

4. Advantages and disadvantages using this method

Compared to the conventional hydrogenation methods as mentioned in other Chapters, the photocatalytic hydrogenation on semiconductor particles has some great advantages: (1) The most important merit is that particular reducing agents (e.g. H2 gas etc.) are not necessary in this method. In most case, the reductants are conventional solvents such as water, alcohols, and amines which concurrently act as hole scavengers. Therefore, this method allows us to avoid both the use of harmful and dangerous chemical reagents and the emission of harmful waste. (2) The reactions mostly proceed under mild conditions, e.g. under ordinary temperature and pressure, and therefore are safety. (3) In the case of TiO2 or other stable metal oxide photocatalysts, the materials are chemically stable, easily removal, and reusable. These three significant advantages imply that this method holds great promise to become an alternative “green” synthetic method.

On the other hand, a disadvantage of this method is to be unsuitable for a large-scale synthesis, because the rate of surface reaction on photocatalysts under irradiation are limited by electron-hole recombination, smaller surface area, lesser adsorptive and diffusive properties of substrates compared to those of the conventional catalysis. Therefore, up to now, scaling-up of the semiconductor photocatalysis has been successfully applied only to wastewater treatment, in which solar photocatalytic degradation of water contaminants, persistent toxic compounds and cyanide etc., is carried out on low concentration of the contaminants [70, 71]. In contrast, highly chemoselective and stereoselective hydrogenations have been made on a pilot plant scale and even fewer are available commercially on a multi kg scale by the conventional heterogeneous or homogeneous hydrogenation methods [72].

Advertisement

5. Concluding remarks and future directions

The photocatalytic hydrogenation on semiconductor particles has developed as the highly efficient and selective reaction during the past three decades. For examples, the selective hydrogenations were reported for the reactions of CH3C≡CH to CH3CH=CH2 on Pt/TiO2 (rutile) [17], the aromatic ketones to the corresponding secondary alcohols on the P25 TiO2 [37], the aliphatic ketones to the corresponding alcohols on the ZnS nano-crystallite [39], and nitroaromatics to the corresponding amino-compounds on TiO2 [67-69]. In addition, the selective formation of methanol from CO2 [53-56] and NH3 from NO3- [65, 66] has been received much attention to solve the environmental issues. Recent progress in the reductive hydrogenation has been directed toward the development of visible light response photocatalysts to utilize solar energy effectively [46-52, 65, 68, 69].

One of the most significant features of semiconductor photocatalysis is that we can utilize both oxidation and reduction in one-pot processes as mentioned in the deaminocondensation reactions [27-30]. The combination of redox reactions can afford several unique reactions, which are not achieved by conventional reaction techniques. Thus, the semiconductor photocatalysis is one of promising methods in fine chemical synthesis for high value pharmaceuticals etc. In order to achieve this method, development of highly stereoselective photocatalysis will be indispensable, although less known about such semiconductor photocatalysts so far. Therefore, particular attention should be directed toward the development of new enantioselective semiconductor catalyzed reactions in the future.

Acknowledgement

The work in ref. [37] (summarized in Table 3) was supported by a Grant-in-Aid for Scientific Research (no. 21590052 and no. 19590005) from the Japan Society for the Promotion of Science, a grant from Prof. Nishihara (HUHS), and Astellas Foundation for Research on Metabolic Disorders.

References

  1. 1. FujishimaA.HondaK.1972Electrochemical Photolysis of Water at a Semiconductor Electrode.Nature2383738
  2. 2. KudoA.MisekiY.2009Heterogeneous Photocatalyst Materials for Water SplittingChem. Soc. Rev. 38253278
  3. 3. FujishimaA.ZhangX.TrykD. A.2007Heterogeneous Photocatalysis: From Water Photolysis to Applications in Environmental CleanupInt. J. Hydrogen Energy 3226642672
  4. 4. HoffmannM. R.MartinS. T.ChoiW.BahnemannD. W.1995Environmental Applications of Semiconductor PhotocatalysisChem. Rev. 956996
  5. 5. FujishimaA.RaoT. N.TrykD. T.2000Titanium Dioxide PhotocatalysisJ. Photochem. Photobiol. C. Photochem. Rev. 1121
  6. 6. SakataT.1985Photocatalysis of Irradiated Semiconductor Surfaces: Its Application to Water Splitting and Some Organic Reactions.J. Photochem. 29205215
  7. 7. Fox M.A1987Selective Formation of Organic Compounds by Photoelectrosynthesis at Semiconductor Particles. Top. Curr. Chem. 1427299
  8. 8. KischH.2001Semiconductor Photocatalysis for Organic SynthesisAdv. Photochem. 2693143and references therein.
  9. 9. PalmisanoG.AugugliaroV.PagliaroM.PalmisanoL.2007Photocatalysis: a Promising Route for 21st Centyry Organic Chemistry. Chem. Commun. 34253437and references therein.
  10. 10. ShiraishiY.HiraiT.2008Selective Organic Transformation on Titanium Oxide-Based Photocatalysts.J. Photochem. Photobiol. C: Photochem. Rev. 9157170and references therein.
  11. 11. PalmisanoG.García-LópezE.MarcíG.ToddoV.YurdakalS.AugugliaroV.PalmisanoL.2010Advances in Selective Conversion by Heterogeneous Photocatalysis.Chem. Commun. 4670747089and references therein.
  12. 12. Boonstra A.H, Mutsaers C.A.H.A1975Photohydrogenation of Ethyne and Ethene on the Surface of Titanium DioxideJ. Phys. Chem. 7920252027
  13. 13. RedmondG.FitsmauriceD.1993Spectroscopic Determination of Flatband Potentials for Polycrystalline TiO2 Electrodes in Nonaqueous Solvents.9714261430
  14. 14. YunC.AnpoM.KodamaS.KubokawaY.1980U.V. Irradiation-Induced Fission of a C=C od C≡C Bond Adsorbed TiO2. J. Chem. Soc., Chem. Commun.609.
  15. 15. AnpoM.AikawaN.KodamaS.KubokawaY.1984Photocatalytic Hydrogenation of Alkynes and Alkenes with Water over TiO2: Hydrogenation Accompanied by Bond Fission. J. Phys. Chem. 8825692572
  16. 16. AnpoM.AikawaN.KubokawaY.1984Photocatalytic Hydrogenation of Alkynes and Alkenes with Water over TiO2: Pt-Loading Effect on the Primary Processes. J. Phys. Chem. 8839984000
  17. 17. AnpoM.ShimaT.KodamaS.KubokawaY.1987Photocatalytic Hydrogenation of CH3CCH with H2O on Small-Particle TiO2: Size Quantization Effects and Reaction Intermediates. J. Phys. Chem. 9143054310
  18. 18. YamatakaH.SetoN.IchiharaJ.HanafusaT.TerataniS.1985Reduction of C-C Multiple Bonds Using an Illuminated Semiconductor CatalystJ. Chem. Soc., Chem. Commun. 788789
  19. 19. BabaR.NakabayashiS.FujishimaA.HondaK.1987Photocatalytic Hydrogenation of Ethylene on the Bimetal-Deposited Semiconductor PowdersJ. Am. Chem. Soc. 10922732277
  20. 20. KodamaS.NakayaH.AnpoM.KubokawaY.1985A Common Factor Determining the Features of the Photocatalytic Hydrogenation and Isomerization of Alkenes over Ti-Si Oxides.Bull. Chem. Soc. Jpn 5836453646
  21. 21. AnpoM.NakayaH.KodamaS.KubokawaY.DomenK.OnishiT.1986Photocatalysis over Binary Metal Oxides: Enhancement of the Photocatalytic Activity of TiO2 in Titanium-Silicon Oxides. J. Phys. Chem. 9016331636
  22. 22. YamashitaH.KawasakiS.IchihashiY.TakeuchiM.HaradaM.AnpoM.LouisC.CheM.1998Characterization of Ti/Si Binary Oxides Prepared by the Sol-Gel Method and Their Photocatalytic Properties: The Hydrogenation and Hydrogenolysis of CH3CCH with H2OKorean J. Chem. Eng. 15491495
  23. 23. LinL.KuntzR. R.1992Photocatalytic Hydrogenation of Acetylene by Molybdenium-Sulfur Complex Supported on TiO2. Langmuir 8870875
  24. 24. Kuntz R.R1997Comparative Study of Mo2OxSy(cys)22- Complexes as Catalysts for Electron Transfer from Irradiated Colloidal TiO2 to Acetylene. Langmuir 1315711576
  25. 25. ShiragamiT.AnkyuH.FukamiS.PacC.YanagidaS.MoriH.FujitaH.1992Semiconductor Photocatalysis: Visible Light Induced Photoreduction of Aromatic Ketones and Electron-deficient Alkenes Catalysed by Quantised Cadmium SulfideJ. Chem. Soc., Faraday Trans. 8810551061
  26. 26. ShiragamiT.FukamiS.WadaY.YanagidaS.1993Semiconductor Photocatalysis: Effect of Light Intensity on Nanoscale CdS-Catalyzed Photolysis of Organic SubstratesJ. Phys. Chem. 971288212887
  27. 27. NishimotoS.OhtaniB.YoshikawaT.KagiyaT. (1983) Photocatalytic Conversion of Primary Amines to Secondary Amines and Cyclization of Polymethylene-α,ω-diamines by an Aqueous Suspension of TiO2/Pt. J. Am. Chem. Soc. 105 71807182 .
  28. 28. OhtaniB.GotoY.NishimotoS.InuiT.1996Photocatalytic Transfer Hydrogenation of Schiff Bases with Propan-2-ol by Suspended Semiconductor Particles Loaded with Platinum DepositsJ. Chem. Soc. Faraday Trans. 9242914295
  29. 29. PalB.IkedaS.KominamiH.KeraY.OhtaniB.2003Photocatalytic Redox-Combined Synthesis of L-Pipecolinic Acid from L-Lysine by Suspended Titania Particle: Effect of Noble Metal Loading on the Selectivity and Optical Purity of the Product.J. Catal. 217152159
  30. 30. OhtaniB.PalB.IkedaS.2003Photocatalytic Organic Synthesis: Selective Cyclization of Amino Acids in Aqueous Suspensions.Catal. Surv. Asia 7165176and references therein.
  31. 31. KünnethR.FeldmerC.KnochF.KischH.1995Semiconductor-Catalyzed Photoaddition of Olefins and Enol Ethers to 1,2-Diazenes: A New Route to Allylhydrazines. Chem. Eur. J. 1441448
  32. 32. WarrierM.LoM. K. F.MonbouquetteH.Garcia-GaribayM. A.2004Photocatalytic Reduction of Aromatic Azides to Amines Using CdS and CdSe NanoparticlesPhotochem. Photobiol. Sci. 3859863
  33. 33. CuendetP.GrätzelM.1987Direct Photoconversion of Pyruvate to Lactate in Aqueous TiO2 Dispersions. J. Phys. Chem. 91654657
  34. 34. Joyce-PrudenC.ProssJ. K.LiY.1992Photoinduced Reduction of Aldehydes on Titanium DioxideJ. Org. Chem. 5750875091
  35. 35. MatsushitaY.KumadaS.WakabayashiK.SakedaK.IchimuraT.2006Photocatalytic Reduction in Microreactors. Chem. Lett. 35410411
  36. 36. MatsushitaY.OhbaN.KumadaS.SakedaK.SuzukiT.IchimuraT.2008Photocatalytic Reactions in Microreactors. 135S: S303S308.
  37. 37. KohtaniS.YoshiokaE.SaitoK.KudoA.MiyabeH.2010Photocatalytic Hydrogenation of Acetophenone Derivatives and Diaryl Ketnoes on Polycrystalline Titanium Dioxide. Catal. Commun. 1110491053
  38. 38. YanagidaS.IshimaruY.MiyakeY.ShiragamiT.PacC.HashimotoK.SatakaT.1989Semiconductor Photocatalysis. ZnS-Catalyzed Photoreduction of Aldehydes and Related DerivativesTwo-Electron-Transfer Reduction and Relationship with Spectroscopic Properties. J. Phys. Chem. 9325762582
  39. 39. YanagidaS.YoshiyaM.ShiragamiT.PacC.MoriH.FujitaH.1990Semiconductor Photocatalysis. Quantitative Photoreduction of Aliphatic Ketones to Alcohols Using Defect-Free ZnS Quantum Crystallites. J. Phys. Chem. 9431043111
  40. 40. YanagidaS.OgataT.ShindoA.HosokawaH.MoriH.SakataT.WadaY.1995Semiconductor Photocatalysis: Size Control of Surface-Capped CdS Nanocrystallites and the Quantum Size Effect in Their PhotocatalysisBull. Chem. Soc. Jpn. 68752758
  41. 41. Park J.W, Hong M.J, Park K.K (2001) Photochemical Reduction of 1,2-Diketones in the Presence of TiO2. Bull. Korean Chem. Soc. 22 12131216 .
  42. 42. Park J.W, Kim E.K, Park K.K2002Photochemical Reduction of Benzil and Benzoin in the Presence of Triethylamine and TiO2 Photocatalyst.Bull. Korean Chem. Soc. 2312291234
  43. 43. ArestaM.2010Carbon Dioxide- Utilization Options to Reduce Its Accumulation in the Atmosphere. In: Aresta M, editor. Carbon Dioxide as Chemical Feedstock. Weinheim: WILEY-VCH 113
  44. 44. InoueT.FujishimaA.KonishiS.HondaK.1979Photoelectrocatalytic Reduction of Carbon Dioxide in Aqueous Suspension of Semiconductor Powders.Nature277637638
  45. 45. HengleinA.GutiérrezM.1983Photochemistry of Colloidal Metal Sulfides. 5. Fluorescence and Chemical Reactions of ZnS and ZnS/CdS Co-Colloids. Ber. Bunsenges. Phys. Chem. 87 852858 .
  46. 46. Pan P-W, Chen Y-W2007Photocatalytic Reduction of Carbon Dioxide on NiO/InTaO4 under Visible Light IrradiationCatal. Commun. 815461549
  47. 47. JiaL.LiJ.FangW.2009Enhanced Visible-light Active C and Fe Co-Doped LaCoO3 for Reduction of Carbon Dioxide. Catal. Commun. 118790
  48. 48. LiuY.HuangB.DaiY.ZhangX.QinX.JiangM.Whangbo-HM.2009Selective Ethanol Formation from Photocatalytic Reduction of Carbon Dioxide in Water with BiVO4 PhotocatalystCatal. Commun. 11210213
  49. 49. YanS. C.OuyangS. X.GaoJ.YangM.FengJ. Y.FanX. X.WanL. J.LiZ. S.YeJ. H.ZhouY.ZouZ. G.2010A Room-Temperature Reactive-Temperature Route to Mesoporous ZnGa2O4 with Improved Photocatalytic Activity in Reduction of CO2. Angew. Chem. Int. Ed. 4964006404
  50. 50. LiuQ.ZhouY.KouJ.ChenX.TianZ.GaoJ.YanS.ZouZ.2010High-Yield Synthesis of Ultralong and Ultrathin Zn2GeO4 Nanoribbons toward Improved Photocatalytic Reduction of CO2 into Renewable Hydrocarbon Fuel.J. Am. Chem. Soc. 1321438514387
  51. 51. ZhangN.OuyangS.LiP.ZhangY.XiG.KakoT.YeJ. (2011) Ion-Exchange Synthesis of a Micro/Mesoporous Zn2GeO4 Photocatalyst at Room Temperature for Photoreduction of CO2. Chem. Commun. 47 20412043 .
  52. 52. ZhouY.TianZ.ZhaoZ.LiuQ.KouJ.ChenX.GaoJ.YanS.ZouZ. (2011) High-Yield Synthesis of Ultrathin and Uniform Bi2WO6 Square Nanoplates Benefitting from Photocatalytic Reduction of CO2 into Renewable Hydrocarbon Fuel under Visible Light. Appl. Mater. Interfaces 3 35943601 .
  53. 53. AnpoM.YamashitaH.IchihashiY.FujiiY.HondaM.1997Photocatalytic Reduction of CO2 with H2O on Titanium Oxides Anchored within Micropores of Zeolites: Effects of the Structure of the Active Sites and the Addition of Pt. J. Phys. Chem. B 10126322636
  54. 54. AnpoM.Kim-HT.MatsuokaM.2009The Design of Ti-, V-, Cr- Oxide Single-Site Catalysts within Zeolite Frameworks and their Photocatalytic Reactivity for the Decomposition of Undesirable Molecules- The Role of their Excited States and Reacition Mechanisms. Catal. Today 142114124
  55. 55. ShioyaY.IkeueK.OgawaM.AnpoM.2003Synthesis of Transparent Ti-Containing Mesoporous Silica Thin Film Materials and their Unique Photocatalytic Activity for the Reduction of CO2 with H2OAppl. Catal. A: General 254251259
  56. 56. MatsumotoY.ObataM.HomboJ.1994Photocatalytic Reduction of Carbon Dioxideon p-Type CaFe2O4 Powder. J. Phys. Chem. 9829502951
  57. 57. BurtT. P.HowdenN. J. K.WorrallF.WhelanM. J.2010Long-Term Monitoring of River Water Nitrate: How Much Data Do We Need? J. Environ. Monit. 127179
  58. 58. KudoA.DomenK.MaruyaK.OnishiT.1987Photocatalytic Reduction of 3-to Form NH3 over Pt-TiO2.Chem. Lett. 1019-1022.
  59. 59. RanjitK. T.ViswanathanB.1997Photocatalytic Reduction of Nitrite and Nitrate Ions over Doped TiO2 CatalystsJ. Photochem. Photobiol. A Chem. 107215220
  60. 60. LiY.WasgestianF.1998Photocatalytic Reduction of Nitrate Ions on TiO2 by Oxalic Acid.J. Photochem. Photobiol. A Chem. 112255259
  61. 61. KudoA.DomenK.MaruyaK.OnishiT.1992Reduction of Nitrate Ions into Nitrite and Ammonia over Some PhotocatalystsJ. Catal. 135300303
  62. 62. KatoH.KudoA.2002Photocatalytic Reduction of Nitrate Ions over Tantalate Photocatalysts428332838
  63. 63. RanjitK. T.KrishnamoorthyR.ViswanathanB.1994Photocatalytic Reduction of Nitrite and Nitrate on ZnSJ. Photochem. Photobiol. A Chem. 815558
  64. 64. KudoA.SekizawaM.2000Photocatalytic H2 Evolution under Visible Light Irradiation on Ni-Doped ZnS Photocatalyst.Chem. Commun. 13711372
  65. 65. KudoA.HamanoiO.2002Reduction of Nitrate and Nitrite Ions over Ni-ZnS Photocatalyst under Visible Light Irradiation in the Presence of a Sacrificial ReagentChem. Lett. 838839
  66. 66. YamauchiM.AbeR.TsukudaT.KatoK.TanakaM.2011Highly Selective Ammonia Synthesis from Nitrate with Photocatalytically Generated Hydrogen on CuPd/TiO2. J. Am. Chem. Soc. 13311501152
  67. 67. MahdaviF.BrutonT. C.LiY.1993Photoinduced Reduction of Nitro Compounds on Semiconductor Particles. J. Org. Chem. 58744746
  68. 68. WangH.YanJ.ChangW.ZhangZ.2009Practical Synthesis of Aromatic Amines by Photocatalytic Reduction of Aromatic Nitro Compounds on Nanoparticles N-Doped TiO2. Catal. Commun. 10989994
  69. 69. FüldnerS.MildR.SiegmundH. I.SchroederJ. A.GruberM.KönigB.2010Green-Light Photocatalytic Reduction Using Dye-Sensitized TiO2 and Transition Metal Nanoparticles. Green Chem. 12400406
  70. 70. MalatoS.BlancoJ.VidalA.RichterC.2002Photocatalysis with Solar Energy at a Pilot-Plant Scale: an OverviewAppl. Catal. B: Environ. 37115
  71. 71. MarugánJ.GriekenR.CassanoA. E.AlfanoO. M. (2009) Scaling-up of Slurry Reactors for the Photocatalytic Oxidation of Cyanide with TiO2 and Silica-Supported TiO2 Suspensions. Catal. Today 144 8793 .
  72. 72. Blaser H-U2010The Development and Application of Industrially Viable Catalysts for the Selective Hydrogenation of Complex MoleculesTop. Catal. 539971001

Written By

Shigeru Kohtani, Eito Yoshioka and Hideto Miyabe

Submitted: 19 December 2011 Published: 10 October 2012