Open access peer-reviewed chapter

Recent Progress of Electrocatalysts and Photocatalysts Bearing First Row Transition Metal for Hydrogen Evolution Reaction (HER)

Written By

Shipra Sagar, Ravi K. Kanaparthi, Manish K. Tiwari and Satyen Saha

Submitted: 24 February 2020 Reviewed: 18 May 2020 Published: 23 June 2020

DOI: 10.5772/intechopen.92854

From the Edited Volume

Photophysics, Photochemical and Substitution Reactions - Recent Advances

Edited by Satyen Saha, Ravi Kumar Kanaparthi and Tanja V. Soldatovi?

Chapter metrics overview

1,011 Chapter Downloads

View Full Metrics

Abstract

The design and modification of metal–organic complexes for hydrogen (H2) gas production by water splitting have been intensively investigated over the recent decades. In most reported mechanistic pathways, metal hydride species are considered as crucial intermediates for H2 formation where the metal present at the active site plays an imperative role in the transfer of electron and proton. In the last few decades, much consideration has been done on the development of non-precious metal–organic catalysts that use solar energy to split water into hydrogen (H2) and oxygen (O2) as alternative fossil fuels. This review discussed the design, fabrication, and evaluation of the catalysts for electrocatalytic and photocatalytic hydrogen production. Mechanistic approach is addressed here in order to understand the fundamental design principle and structural properties relationship of electrocatalysts and photocatalysts. Finally, we discuss some challenges and opportunities of research in the near future in this promising area.

Keywords

  • redox-active-ligand
  • first-row transition metals
  • hydrogen evolution
  • catalytic cycle
  • electrocatalyst
  • photocatalyst

1. Introduction

Climate change and increasing energy demand have emphasized research on sustainable energy source [1, 2]. Day-by-day increase of human population and global requirements has compelled researchers to develop new renewable sustainable energy sources in replacement of hydrocarbon deposits [3]. Renewable sources such as solar power, wind, and water, storage of these energies for on-demand utilization, and transportation are the major challenges for researchers. To develop a clean and eco-friendly environment, splitting of water into hydrogen and oxygen is a tremendous way to produce sustainable energy. Hydrogen gas emerged as a green energy fuel due to its high-energy density and zero carbon dioxide (CO2) emission [4, 5, 6]. In this regard, electrocatalytic and photocatalytic H2 generation from water has been considered as one of the most striking approaches [7, 8, 9]. In recent years, a substantial number of artificial photosynthesis have been developed, exploited solar power as electron and proton source to make a clean renewable fuel [10, 11, 12, 13, 14]. Light-induced splitting of water is a suitable process because the production of hydrogen is used as green fuel in future and even used for the synthesis of other chemicals [15, 16, 17, 18].

Literature reports suggested more than 500 billion cubic meters (44.5 million tons) of hydrogen gas is produce yearly worldwide [19, 20]. In the current scenario, steam methane refining, coal gasification, and water electrolysis are the major way for hydrogen production. Nowadays 95% hydrogen gas is produced from steam methane reforming and coal gasification, however only 4% hydrogen from water electrolysis. Steam methane is a high-energy-intensive process maintained at high temperature with the formation of carbon dioxide and carbon monoxide: (i) CH4 + H2O = CO + 3H2 (ii) CO + H2O = CO2 + H2. Hence, it is not an eco-friendly method for hydrogen production. Water electrolysis is the most sustainable and clean approach for hydrogen production because its source is abundant. Since the most suitable way of light-driven energy conversion is water electrolysis, artificial photosynthesis (PS II) has been considered as primary goal to produce electron and proton [21, 22]. Water splitting is a redox reaction in which aqueous protons are reduced into H2 at cathode and water is oxidized to O2 at anode [23]. Both H2 (HER) and O2 (OER) reactions are rigorously coupled, which may lead to the formation of explosive H2/O2 mixtures due to gas crossover [24, 25, 26]. By far, only a few stable metal complexes as catalysts are achieved that can decompose water into H2 and O2 [27, 28, 29, 30, 31]. Water-splitting reactions are split into two half-reactions: water oxidation to O2 evolution and water reduction to H2 production:

H2O2H2+O2 E0=1.23VE1
2H2O4H++4e+O2E0=1.23Vvs.NHEE2
4H++4e2H2E0=0Vvs.NHEE3

The limitation of OER is that it takes place after the successive accumulation of four oxidized electrons and protons in Kok cycle (catalytic cycle of the water oxidation in PS II) that require much higher overpotential input than that of HER [32]. Thermodynamic potential is different for H+/H2 (0 V vs. NHE) and OH-/O2 (1.23 V vs. NHE), and the overall solar energy conversion efficiency is only ∼15% in OER [33]. The hydrogen evolution reaction (HER, 2H+ + 2e- = H2) is the cathodic reaction with the two-electron transfer in one catalytic intermediate and offers the potential to hydrogen production. However, hydrogen production technology requires proficient electrocatalysts and photocatalysts which support two key electrode reactions (OER and HER) at lower overpotentials.

Moreover discussion on the mechanism of HER, H+ adsorption on the hydrogen evolution catalyst surface is the first step, known as Volmer step, followed by Heyrovsky or Tafel steps shown in Figure 1. A suitable HER catalyst always binds H+ very fast and releases the product. Hence, electrochemical hydrogen evolution reaction (HER) facilitates for H2 production on large-scale.

Figure 1.

The inside mechanism of H2 evolution of electrocatalyst in acidic solution.

Afterwards, H2 evolution may occur via two different reaction mechanisms depending on the action of catalyst [34]. Hydronium cation (H3O+) is the proton source in acidic solution, and in alkaline condition H2O is the proton source. In Volmer-Tafel mechanism, two protons absorbed on the catalytic surface can combine to form H-H bond to yield H2. In Heyrovsky reaction route, a second electron and another proton from the solution are transferred to the catalyst surface which reacts with the absorbed H atom and generate H2. This is an electrochemical desorption pathway. Precious metal like Pt-based electrocatalysts is highly reactive for HER and is usually pursuing Volmer-Tafel mechanism. Lately few literatures [35, 36, 37] have been reported on Ni-based electrocatalysts which follows Volmer-Heyrovsky path.

1.1 Fundamental aspects of photocatalytic and electrocatalytic hydrogen production process

Electrocatalytic water splitting is driven by passing the electric current through the water; conversion of electrical energy to chemical energy takes place at electrode through charge transfer process. During this process, water reacts at the anode form O2 and hydrogen (proton) produce at the cathode as we mentioned earlier. Suitable electrocatalysts can maximally reduce the overpotential which is highly desirable for driving a specific electrochemical reaction. However, the process of surface catalytic reactions in electrocatalysis is very similar to photocatalysis [38]. Photocatalytic is a simple water-splitting reaction in which H2 and O2 are produced from water by utilizing the energy of sunlight. Figure 2(a) shows the process of photocatalysis in which a metal catalyst contains chromophores that immersed solar energy and triggered the electron transfer reaction. The most important criteria for the solar-driven water-splitting reaction are electronic band gap matching of the photosensitive material to the redox potential of water [39]. Metal complexes act as chromophore associated with mainly three types of electron transfer: metal center (MC), ligand center (LC), and metal–ligand center transition (MLCT). The MLCT state of the metal complex plays a crucial role in photocatalytic reactions. In octahedral complexes with conjugated ligand system, the highest occupied molecular orbital (HOMO) corresponds to the metal-localized t2g-orbitals, and the lowest unoccupied molecular orbital (LUMO) is associated with anti-bonding π*-orbital localized on the ligands. On the absorption of UV–visible light, an electron is promoted from one of the metal-centered t2g orbitals to a ligand-centered π* orbital, resulting in the MLCT state shown in Figure 2(b). As a result, the redox properties of the metal complexes are dramatically changed. The excited metal complexes behave as better oxidants and better reductants than their electronic ground state and can hold more thermodynamic driving force for the charge transfer reactions. Based on the photo-induced redox potential changes and the long-lived lifetime of the excited state, many metal complexes have been intensively investigated as chromophores for this photocatalytic H2 production purpose [40]. Zou and coworkers have described various photocatalytic systems for H2 production, which exposed that most of the photocatalytic systems suffer photodecomposition and instability [40]. Hence, for long-term use, it is imperative to build up highly proficient H2 generation systems with long lifetimes and high durability. Many reviews have been published on solar H2 evolution systems based on photocatalysts [41, 42, 43, 44].

Figure 2.

(a) Photocatalyst system for water splitting. (b) Molecular orbital diagram for d6 metal complex chromophores.

1.2 Mechanistic pathway

So far, extensive theoretical study has been revealed, the possible mechanistic process of proton reduction to hydrogen evolution through transition metal molecular catalyst. A generalized mechanistic scheme depicting the homolytic and heterolytic path is shown in Figure 3. The homolytic mechanism involves bimetallic route, where a metal hydride species ([Mn+–H]) react with another metal hydride to release one H2 via reductive elimination. Instead, heterolytic is a monometallic pathway, where the metal hydride [Mn+–H] is further reduce and protonated for H2 evolution [45]. Both pathways function simultaneously, two protons and two electrons are delivered to the metal center, and in few cases, the pH, catalytic concentration, and proton source decide the dominant route [46]. During the past decade, a number of review articles emphases on the structural property relationship and mechanistic study [45, 47, 48, 49]. Among all research on catalyzed H2 evolution, the mechanistic investigation on proton reduction catalysis is essential because it can give us a significant idea to design better molecular catalysts in the future [49].

Figure 3.

Proposed homolytic and heterolytic mechanisms for H2 evolution via the formation of a metal hydride (M-H) intermediate.

1.3 Metal and redox-active ligands for HER

Here we start by describing the fundamental concept of metal and organic ligand system which gives a strong influence on the performance of H2 evolution. Transition metal cations with partial filled d-electronic configurations are considered as catalyst. The characteristic feature of this type of catalyst is that the metal ions can exist in higher oxidation state [44]. There are several literatures reported with partially filled d-orbital which show high stability toward water-splitting reactions [50, 51, 52, 53]. However, the most catalytic system requires very high temperature and precious metal at the active site; therefore, it will be a challenge for researchers to develop a photocatalytic and electrocatalytic system at low temperature with low-cost metal [54, 55, 56, 57]. Indeed, first-row transition metal complexes (Co, Ni, Cu, Zn) have been exploited in the last few decades for this purpose. Beyond the reactivity of metal, the redox activity of organic ligands has also received continuous attention. The redox-active ligand works as electron sink in the complexes and maintains the metal in its original oxidation state. Redox-active ligands convey a novel reactivity to the complex by loss or gain of electrons [58]. In addition, the redox-active property of the ligand can also be influenced by the modification of the substituents by σ and π donating ability, π accepting ability, and conjugation [59]. A highly conjugated system such as bpy, porphyries, and ortho-phenylenediamine (opda) having anti-bonding π*-orbital localized on the ligands is considered for hydrogen production due to its multielectron or multiproton pooling ability which is responsible for dramatically changing the potential of redox properties [60, 61]. Substituents attached to redox-active ligands, electron density, and charge on the metal ions also effect the standard electrode potential.

So far, considerable advancement has been done in the field of electrocatalytic and photocatalytic water-splitting reaction for hydrogen production, and several advance review papers have been reported by scientists [40, 51, 62, 63, 64]. However, very limited comprehensive tutorial has published on only first-row transition metal-based catalysts. This chapter describes electrocatalytic as well photocatalytic properties of inorganic catalysts and their structural and mechanistic features. Here we put an effort elucidate the direction of fundamental mechanistic aspects during electrocatalytic and photocatalytic hydrogen (H2) production reaction (HER).

Advertisement

2. Photochemical hydrogen production from a series of 3D transition metal complexes bearing o-phenylenediamine ligand

Masaki Yoshida et al. [65] developed a series of 3D-transition metal complexes with o- phenylenediamine (opda) ligands for hydrogen production due to the following properties: (a) aromatic amine undergoes homolytic N-H bond cleavage by photoexcitation [66] which is applicable for hydrogen production under mild condition, and (b) opda complexes have extensively been obsessed as reversible multielectron or multiproton pooling ability because of its multistep redox processes between o-phenylenediamine(opda), semibenzoquinodiamine (s-bqdi), and o-benzoquinodiimine (bqdi) which is useful for reversible hydrogen production. [M(opda)3], M = Mn2+, Fe2+, Co2+, Ni2+, and Zn+2 complexes (1–5) shown in Table 1 have photochemical hydrogen production ability.

  1. a. Photochemistry of aromatic amine (opda)

  1. b. Redox properties of opda complex

Table 1.

Structure of redox-active ligands, electrocatalysts, and photocatalysts.

In the past M. Yoshida and coworkers proposed [Fe-opda] for photochemical HER mechanism at photoirradiation of λ = 298 nm; ππ* excitation occurred in complex with N-H bond homolysis process, followed by H2 elimination [67]. After this process, the opda ligands in the complex were partially oxidized to bqdi or s-bqdi ligands. This mechanism is based on the deep-rooted photochemical N-H bond activation of aromatic amines. Theoretical study and ultrafast spectroscopic studies of amino benzene support that the photochemical N-H fission occurs by the photoexcitation to higher-lying ππ* level which leads to the formation of the πσ* state [68]. Photochemical mechanism for HER of all complexes is shown in Figure 4.

Figure 4.

Plausible mechanism for photochemical HER with [M-opda] complexes.

All opda-based metal complexes display photochemical HER activities with the formation of almost one equivalent of H2 gas. However, the HER was not observed at all in the dark in all complexes, which suggests that the HER was obsessed by photochemical reaction. Moreover, they observed remarkable decrease in hydrogen evolution reaction, while the ligand is replaced with aromatic amines. This experiment suggested that the photo-induced HER activities of the complexes in this case are weakly dependent on the central metal ion and strongly dependent on the redox-active ligand. Further to check the metal ion dependency, examine the catalytic hydrogen production in the presence of hydroquinone (HQ; 10 equiv) as a sacrificial electron-proton donor. The photochemical H2 production from [M-opda] (7.98 × 10−2 mmol) with HQ (7.98 × 10−1 mmol) in THF (4 mL) under an N2 atmosphere at 20°C for 190 h turns over the number for all the complexes given in Table 2. Difference in TON may be caused by the stability of each complexes.

CatalystsRedox-active organic ligandsCatalytic potential (Ep)SolventTON(H2 mol cat−1)Ref.
[1]ao-Phenylenediamine----THF5.5[65]
[2]o-Phenylenediamine----THF2.9[67]
[3]o-Phenylenediamine----THF0.99[65]
[4]o-Phenylenediamine----THF0.51[65]
[5]o-Phenylenediamine----THF0.73[65]
[6]bDiimine-dioxime−0.68 V vs. Fc+/0H2O/CH3CN300[70]
[7]Diimine-dioxime−0.96 V vs. Fc+/0H2O/CH3CN50[70]
[8]Bis(thiosemicarbazone)−1.7 V vs. Fc/Fc+CH3CN37[74]
[9]Bis(thiosemicarbazone)−1.7 V vs. Fc/Fc+CH3CN73[75]
[10]cDiamine-tripyridine−0.90 V vs. Fc+/0acidic-H2O1.4x 104[70]
[11]dTMPA−1.81 V vs. SCECH3CN/H2O6180[76]
[12]Cl-TMPA−1.72 V vs. SCECH3CN/H2O10,014[76]
[13]Bis(benzenedithiolate)−2.25 V vs. SCECH3CN0[77]
[14]o-Aminobenzenethiolate−1.64 V vs. SCECH3CN6190[77]
[15]o-Aminobenzene−2.03 V vs. SCECH3CN900[77]
[16]2-Mercaptophenolate−1.62 V vs. SCECH3CN5600[77]

Table 2.

Electrochemical data and catalytic efficiency of metal complexes for water-splitting hydrogen evolution reaction.

a = photochemical H2 production from [M-opda 1–5] (7.98 × 10−2 mmol) with HQ (7.98 × 10−1 mmol) in (4 mL) under an N2 atmosphere at 20°C for 190 h in the presence of 4 AMS (irradiation with 300 W Xe lamp, 250–385 nm). b = in light-driven condition of TEA and a cyclometalated iridium-based photosensitizer; CV was done in glassy carbon electrode (100 mVs−1) in the presence of p-cyanoanilinium tetrafluoroborate. c = the controlled potential electrolysis of complex 11 was measured at pH 2.5 under N2 atmosphere at a scan rate of 50 mVs−1in phosphate buffer, over 2 hours using a glassy carbon electrode. d = conditions, 0.1 M n-Bu4NPF6, scan rate 100 mV/s, in Ar-saturated CH3CN/H2O.

2.1 Cobalt diimine-dioxime complexes for HER

V. Artero and coworker synthesized cobalt diimine-dioxime complexes as molecular catalysts for H2 evolution [69, 70]. This synthesized ligand (N2, N2propanediylbis-butan-2-imine-3-oxime) has emerged many years ago through Schiff base condensation of butanedione-monoximeon diamine compounds but not widely used for HER. Cobalt diimine-dioxime catalysts are active for H2 evolution in aqueous solution, both after immobilization on electrode materials and in light-driven homogeneous conditions. The electrocatalytic activity of complexes (6) and (7) for hydrogen evolution was supported gas chromatography and by cyclic voltammetry study illustrated in Figure 5 with the appearance of an irreversible wave. In the case of the BF2-associated complex (7), the electrcatalytic wave which developed toward negative potentials corresponds to the CoII/CoI couple in contrast to the H-bridged complex (6), electrocatalytic H2 evolution occurs at potentials positively shifted with regard to the CoII/CoI couple.

Figure 5.

Cyclic voltammograms of (6) Br2 and (7) Br2 (1 mM, black traces) recorded in CH3CN at a glassy carbon electrode at a speed of 100 mV·s−1 (the figure is reproduced from Ref. [69], with permission from the publisher).

2.2 Electrocatalytic cycle for H2 evolution

In the electrocatalytic cycle, V. Artero and coworker observed that in acetonitrile medium, halide ligands are banished with reduced oxidation state from CoII to CoI. Upon reduction, the coordination in number decreases from six in CoII state to five in CoI state; this characteristic was supported by DFT calculations [71]. In the catalytic cycle (Figure 6), the first step is the transfer of electron and proton by proton-coupled electron transfer (PCET) process. PCET is a chemical reaction that involves the transfer of electron and proton in which the oxidation number changes by CoII to CoI. In the second steps, further electron and proton transfer takes place by PCET process, and the oxidation number changes from CoI to CoII. In the last step, H2 is produced in dihydrogen bond through an intramolecular mechanism. The authors also confirmed cobalt diimine-dioxime catalysts 6 and 7 active for H2 evolution under light-driven conditions in the presence of photosensitizers, associated with Ru, Ir, or Re derivatives. The photocatalytic activity of 6-Br2 was observed in mixed H2O/CH3CN solvent in the presence of TEA and cyclometalated iridium-based photosensitizer [72]. Turnover numbers (TON) determined after continuous 4 h UV–visible light irradiation 6-Br2 showed 300 (H2 mol cat−1), whereas 7-Br2 complex shows only 50 (H2 mol cat−1) as shown in Table 2.

Figure 6.

Possible pathway for catalytic hydrogen evolution, involving PCET processes.

Similarly, cobalt bis(iminopyridine) complex 8 was prepared for electrocatalytic water-splitting reaction [73]. The ligand-centered redox activity was observed during the cyclic voltammetry, suggesting the considerable role of redox-active ligand which is completely involved to stabilize the cobalt metal in higher oxidation state. The two reduction potentials were observed for the CoIII/CoII (quasi reversible) and CoII/CoI (reversible) couples at −0.34 V and −0.86 V (vs. Ag/AgC), respectively. The improved water reduction was attributed to the assimilation of a redox-active-ligand to cobalt which facile reduction of [Co-H]n+ species.

2.3 ZnII and CuII complexes for HER

Grapperhaus et al. [74] recently reported two homogeneous electrocatalysts for H2 production. They derived bis(thiosemicarbazones) ligand from 1,2-diones, considered as a kind of multitalented redox non-innocent system. Tetra-coordinated N2S2 is able to bind with low-valent transition metals centered and formed to stable neutral complexes (9, 10) (Table 1). ZnII complex 9 containing a redox-active moiety diacetyl-bis (N-4 methyl-3-thiosemicarbazide) exhibits the homogeneous catalysis of electro-driven H2 evolution through proton reduction with a maximum of TOF 1170 s−1 in CH3OH and 11,700 s−1 in CH3CN at an overpotential of 0.756 V and 1.074 V, CH3COOH used as proton source. Bulk electrolysis showed that the TON of H2 evolution of ZnII is 37 in over 2.5 h experiments. To make this more comparative, Grapperhaus and coworker synthesized another CuII complex with similar ligand diacetyl-bis(N-4-methyl-3-thiosemicarbazide) and examined H2 evolution reaction. CuII complex 10 exhibits a maximum TOF of 10,000 s−1 in CH3CN and 5100 s−1 in DMF at an overpotential of 0.80 and 0.76 V, respectively. Controlled potential electrolysis confirmed CuII complex act as an excellent electrocatalyst to produce H2 with a minimum faradic efficiency of 81% and TON as high as 73 during experiment over 23 h. They examined HER mechanism of complex 10 through DFT computational studies. In the proposed mechanism, initially the protonation occurs at the hydrazino nitrogen ligand. This was followed by an electrochemical step as a formation of reduced CuI species [CuI(HL)]. The second protonolysis occurs at the opposite hydrazine nitrogen of the ligand to yield [CuI(H2L)]+. Further one-electron reduction of [CuI(H2L)]+ leads to the formation of the H2 evolution illustrated in Figure 7. Hence, here it is worth to mention that the identity of metal ions at the active site affects the HER mechanism.

Figure 7.

Plausible mechanism for proton reduction in complex [CuL].

Professor Wang and group proposed [75] a significant homogeneous mononuclear copper electrocatalyst for H2 production attributed to diamine-tripyridine ligand; complex 11 attains trigonal bipyramidal geometry. According to the author, this ionic copper complex [Cu(bztpen)]2+ with a five coordinating nitrogen atom shows a Jahn-Teller effect. Electrochemical and spectroscopic studies supported that the H2 generation reaction takes place by two successive proton-coupled reduction processes. On the experimental observations of DPV, CV, UV–vis, and 1H-NMR spectroscopic study, the authors proposed two possible pathways: path(A), protonation takes place at the CuI centered in the first step (Eqs. (4) and (5)), and path(B), protonation occurs at one of the nitrogen atoms of the ligand (Eqs. (6), (7), and (8)). H2 generation reaction in path B takes place by two successive proton-coupled reduction processes. Protonation occurs at the ligand centered in the first reduction step, followed by the CuI centered in the second step. This provides [(bztpenH)CuII(H)]2+copper hydride species, which release H2 and regenerate CuII catalyst.

Path A:

bztpenCuII2++e+H+0.03VbztpenCuIIIH2+E4
bztpenCuIIIH2++e+H+0.85VbztpenCu(II2++H2E5

Path B:

bztpenCuII2++e+H+0.03VbztpenHCuI2+E6
bztpenHCuI2++e+H+0.85VbztpenHCuIIH2+E7
bztpenHCuIIH2+bztpenCuII2++H2E8

According to the author’s studies on the mechanism of this process, the controlled potential electrolysis of complex 11 was measured at pH 2.5 in phosphate buffer at −0.90 V, over 2 h in a glassy carbon electrode. TON 1.4 × 104 mol H2 (mol cat−1) cm−2 was calculated on a faradic efficiency of approximately 96%, which corresponds to a TOF of 2.0 molH2 (mol cat−1)s−1cm−2of [(bztpen)Cu](BF4)2.

Moreover, Wang et al. [76] fabricated and examined two Cu complexes with TMPA = tris(2-pyridyl)methylamine and Cl-TMPA 1-(6-chloropyridin-2-yl)methyl-N,N-bis(pyridin-2-ylmethyl)methaneamine for photocatalytic H2 evolution behavior. They observed both in Cu(II) complexes [Cu(TMPA)Cl]Cl (12) and [Cu(Cl-TMPA)Cl2](13) that (13) is far efficient for photocatalytic H2 production than (12), due to the presence of more labile Cl ligand with longer Cu-Cl bond length and a dangling Cl-substituted pyridyl unit in the second coordination sphere, which both contribute to a higher photocatalytic activity of complex (13). TMPA acts as a tetradentate ligand and coordinate with Cu(II) in a distorted trigonal manner; Cl-TMPA acts as a tridentate ligand coordinate to Cu(II) with two chloride ions in a distorted square pyramidal manner, leaving one Cl-substituted pyridyl group in the second coordination sphere structure which is given in Table 1. ESI-Ms data favor the formation of Cu-hydride intermediate for hydrogen evolution. The authors investigated the photocatalytic H2 production activities in the presence of a multicomponent [Ir(ppy)2(dtbpy)]Cl (ppy = 2-phenylpyridine, dtbpy = 4,4′-di-tert-butyl-2,2′-bipyridine) and triethylamine (TEA) photosystems as sacrificial reductant (SR) under optimal condition upon 6 h of irradiation of UV–visible light, the turnover number (TON) of which is calculated as 6108 for complex (12) and 10,014 for complex (13).

Based on the control potential electrolysis experimental data, the authors proposed photocatalytic hydrogen evolution mechanism. In the first step, excited PS system takes out one electron from TEA and donates to CuII center of complex (13). The protonated Cl-substituted pyridyl unit accepts that electron and kicks out the Cu-Cl center for the dissociation of Cl ligand which is substituted at the apical position which is more labile (longest bond length Cu-Cl). After that, the CuI species accept one e and one H+ from the reduced Cl-substituted pyridinium moiety; CuII-H center is formed as a key intermediate which lead to H2 evolution. This executive mechanism provides us guidelines to design more efficient Cu-based catalysts for WRCs in the near future (Figure 8).

Figure 8.

Photocatalytic H2 evolution mechanism for complex [Cu(Cl-TMPA)Cl2] 12.

2.4 Ni electrocatalysts for HER

Professor Richard Eisenberg and coworker [77] synthesized a sequence of nickel bis(chelate) complexes; all complexes attained square planar geometry and examined photocatalytic as well as electrocatalytic behavior for hydrogen evolution. Fluorescein (Fl) as the photosensitizer along with triethanolamine (TEOA) as the sacrificial electron donor was used in water under basic medium (pH = 9.8). Bis(chelate) complexes (14–17) contain bdt(bisdithiolate) and their derivatives having S, O, or N as donors for coordination to the Ni center. The photochemical study reveals that only complexes 15 and 17 exhibited similar activity of hydrogen production in terms of TON (6000 mol of H2 per mole of catalyst in 100 h). In contrast, the TON for the sample complex 16 was much lower (∼900). Complex 14 was observed to be inactive for hydrogen generation. To understand the better significant result, the authors pursued electrochemical experiments under a fixed potential to examine the catalytic activity of complexes. In the absence of acid, both monoanionic complexes 14 and 17 exhibit only one reversible redox couple at −0.56 and −0.45 V, respectively, which is attributed to the [NiL2]/[NiL2]2−ligand-based reductions. The neutral complexes 15 and 16 exhibit two reversible reduction waves at −0.11 and −0.94 V and at −0.89 and −1.53 V, respectively. However, on increasing the concentration of acetic acid to these solutions, more negative catalytic waves were observed in cyclic voltammetry study. Complex 14 shows much more negative reduction potential (−2.25) which suggested this system is inactive for H2 generation (as we discussed earlier). As is the case for complex 14, complex 16 exhibits a catalytic wave potential at −2.03 V which showed poor activity for H2 generation as a consequence of not favorable electron transfer step from Fl to the catalyst. Moreover, complexes 15 and 17 show activity for light-driven H2 generation; these two complexes display electrocatalytic wave potential at significantly less negative at −1.64 and −1.62 V, respectively.

Advertisement

3. Concluding remarks and future scope

Here we present a recent development of molecular catalysts toward clean and renewable fuels using earth-abundant metals. We have highlighted a series of Co-, Ni-, Cu-, Zn-based complexes for HER. We have recapitulated the fundamental principles of hydrogen and oxygen evolution reactions with molecular complexes. The designing and fabrication of the molecular complexes with redox-active ligands have been discussed in details; HER activity of the complexes strongly dependent on redox-active ligands as well the central metal ions are discussed in detail. A mechanistic approach and transfer of electron and proton during the homogeneous electrocatalyst and photocatalysts cycle are given in point. Although reasonable progress has been made in the development of metal complexes based electrocatalysts and chromospheres for photocatalytic hydrogen production, still several issues exist which need further improvement: (i) some photocatalytic systems suffer from low activities and short life times which is manifested in the instability of catalytic systems and so concern on the systems with modest water splitting activity and poor stability of the complexes. (ii) most of the complexes are not soluble in water leading to the use of organic solvent or mixture of organic- water solvent. From the future prospective, it is required to develop redox-active ligands with substituted functional group to increase the solubility of complexes in water. More experimental, spectroscopic, magnetic, and theoretical investigations is still needed to be carried out in order to understand the ligand- and metal-centered electron transfer processes. (iii) In addition, the overpotential requirements for most of the organic ligands are still very high, a chelating ligands giving much lower thermodynamic potentials and much smaller oxidation potential that should be utilized in future. (iv) In the regard of future growth in this field, with the need to design molecular complexes that can be immobilized on the surface of the electrode, for this purpose addition of suitable functional group in the ligand is necessity. These complexes can also be supported by the development of surface of the solid photocatalyst, like TiO2, BiVO4, etc. to demonstrate efficient photoelectrochemical cell.

Advertisement

Acknowledgments

The authors acknowledges the helps received from Ms. Priyadarshini Sahu and Mr. Abhineet Verma during the manuscript preparation.

References

  1. 1. Lewis NS, Nocera DG. Powering the planet: Chemical challenges in solar energy utilization. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:15729-15735
  2. 2. Kim D, Sakimoto KK, Hong D, Yang P. Artificial photosynthesis for sustainable fuel and chemical production. Angewandte Chemie, International Edition. 2015;54:3259-3266
  3. 3. Esswein AJ, Nocera DG. Hydrogen production by molecular photocatalysis. Chemical Reviews. 2007;107:4022-4047
  4. 4. Tollefson J. Green carbon as a bridge to renewable energy. Nature. 2010;464:1262-1264
  5. 5. Pagliaro M, Konstandopoulos AG, Ciriminna R, Palmisano G. Solar hydrogen: Fuel of the near future. Energy & Environmental Science. 2010;3:279-287
  6. 6. Armaroli N, Balzani V. The hydrogen issue. ChemSusChem. 2011;4:21-36
  7. 7. Nocera DG. Solar fuels and solar chemicals industry. Accounts of Chemical Research. 2017;50:616-619
  8. 8. Stamenkovic VR, Strmcnik D, Lopes PP, Markovic NM. Energy and fuels from electrochemical interfaces. Nature Materials. 2017;16:57-69
  9. 9. Hwang J, Rao RR, Giordano L, Katayama Y, Yu Y, Shao-Horn Y. Perovskites in catalysis and electrocatalysis. Science. 2017;358:751-756
  10. 10. Andreiadis ES, Chavarot-Kerlidou M, Fontecave M, Artero V. Artificial photosynthesis: From molecular catalysts for light-driven water splitting to photoelectrochemical cells. Photochemistry and Photobiology. 2011;87:946-964
  11. 11. Joya KS, Joya YF, Ocakoglu K, van de Krol R. Water-splitting catalysis and solar fuel devices: Artificial leaves on the move. Angewandte Chemie, International Edition. 2013;52:10426-10437
  12. 12. Berardi S, Drouet S, Francis L, Gimbert-Surinach C, Guttentag M, Richmond C, et al. Visible light driven water splitting in a molecular device with unprecedentedly high photocurrent density. Chemical Society Reviews. 2014;43:7501-7519
  13. 13. Song W, Chen Z, Brennaman MK, Concepcion JJ, Patrocinio AOT, Iha NYM, et al. Making solar fuels by artificial photosynthesis. Pure and Applied Chemistry. 2011;83:749-768
  14. 14. Gust D, Moore TA, Moore AL. Solar fuels via artificial photosynthesis. Accounts of Chemical Research. 2009;42:1890-1898
  15. 15. Zou Z, Ye J, Sayama K, Arakawa H. Direct splitting of water under visible light irradiation with an oxide semiconductor photocatalyst. Nature. 2001;414:625-627
  16. 16. Hara M, Kondo T, Komoda M, Ikeda S, Shinohara K, Tanaka A, et al. Cu2O as a photocatalyst for overall water splitting under visible light irradiation. Chemical Communications. 1998;3:357-358
  17. 17. Wang X, Xu Q, Li M, Shen S, Wang X, Wang Y, et al. Photocatalytic overall water splitting promoted by an α–β phase junction on Ga2O3. Angewandte Chemie, International Edition. 2012;51:13089-13092
  18. 18. Liao L, Zhang Q, Su Z, Zhao Z, Wang Y, Li Y, et al. Efficient solar water-splitting using a nanocrystalline Co-O photocatalyst. Nature Nanotechnology. 2014;9:69-73
  19. 19. Balat M. Potential importance of hydrogen as a future solution to environmental and transportation problems. International Journal of Hydrogen Energy. 2008;33:4013-4029
  20. 20. Wang M, Wang Z, Gong X, Guo Z. The intensification technologies to water electrolysis for hydrogen production, a review. Renewable and Sustainable Energy Reviews. 2014;29:573-588
  21. 21. Xu Y, Kraft M, Xu R. Metal-free carbonaceous electrocatalysts and photocatalysts for water splitting. Chemical Society Reviews. 2016;45:3039-3052
  22. 22. Oshima T, Lu D, Ishitani O, Maeda K. Intercalation of highly dispersed metal nanoclusters into a layered metal oxide for photocatalytic overall water splitting. Angewandte Chemie, International Edition. 2015;54:2698-2702
  23. 23. Chen X, Shen S, Guo L, Mao SS. Semiconductor-based photocatalytic hydrogen generation. Chemical Reviews. 2010;110:6503-6570
  24. 24. Rausch B, Symes MD, Chisholm G, Cronin L. Decoupled catalytic hydrogen evolution from a molecular metal oxide redox mediator in water splitting. Science. 2014;345:1326-1330
  25. 25. Symes MD, Cronin L. Decoupling hydrogen and oxygen evolution during electrolytic water splitting using an electron-coupled-proton buffer. Nature Chemistry. 2013;5:403-409
  26. 26. Rausch B, Symes MD, Cronin L. A bio-inspired, small molecule electron-coupled proton buffer for decoupling the half-reactions of electrolytic water splitting. Journal of the American Chemical Society. 2013;135:13656-13659
  27. 27. Zou Z, Ye J, Sayama K, Arakawa H. Direct splitting of water under visible light irradiation with an oxide semiconductor photocatalyst. Nature. 2001;414:625-627
  28. 28. Zhang P, Wang M, Li C, Li X, Dong J, Sun L. Photochemical H2 production with noble-metal-free molecular devices comprising a porphyrin photosensitizer and a cobaloxime catalyst. Chemical Communications. 2010;46:8806-8809
  29. 29. Bartelmess J, Weare WW, Sommer RD. Synthesis, characterization and structural investigation of novel meso-pyridyl BODIPY–cobaloxime complexes. Dalton Transactions. 2013;42:14883-14891
  30. 30. Yuan YJ, Tu JR, Lu HW, Yu ZT, Fan XX, Zou ZG. Neutral nickel(ii) phthalocyanine as a stable catalyst for visible-light-driven hydrogen evolution from water. Dalton Transactions. 2016;45:1359-1363
  31. 31. Karunadasa HI, Chang CJ, Long JR. A molecular molybdenum-oxo catalyst for generating hydrogen from water. Nature. 2010;464:1329-1333
  32. 32. Kok B, Forbush B, McGloin M. Cooperation of charges in photosynthetic O2 evolution - I, a linear four step mechanism. Photochemistry and Photobiology. 1970;11:457-475
  33. 33. Askerka M, Brudvig GW, Batista VS. The O2-evolving complex of photosystem II: recent insights from quantum mechanics/molecular mechanics (QM/MM), extended x-ray absorption fine structure (EXAFS), and femtosecond x-ray crystallography data. Accounts of Chemical Research. 2017;50:41-48
  34. 34. Morales-Guio CG, Stern L, Hu X. Nanostructured hydro treating catalysts for electrochemical hydrogen evolution. Chemical Society Reviews. 2014;43:6555-6569
  35. 35. Laursen AB, Patraju KR, Whitaker MJ, Retuerto M, Sarkar T, Yao N, et al. Nanocrystalline Ni5P4: A hydrogen evolution electrocatalyst of exceptional efficiency in both alkaline and acidic media. Energy & Environmental Science. 2015;8:1027-1034
  36. 36. Huang Z, Chen Z, Chen Z, Lv C, Meng H, Zhang C. Ni12P5 nanoparticles as an efficient catalyst for hydrogen generation via electrolysis and photoelectrolysis. ACS Nano. 2014;8:8121-8129
  37. 37. Feng L, Vrubel H, Bensimon M, Hu X. Easily-prepared dinickel phosphide (Ni2P) nanoparticles as an efficient and robust electrocatalyst for hydrogen evolution. Physical Chemistry Chemical Physics. 2014;16:5917-5921
  38. 38. Li R, Li C. Chapter one-Photocatalytic water splitting on semiconductor-based Photocatalysts. Advances in Catalysis. 2017;60:1-57
  39. 39. Vozniuk O, Tanchoux N, Millet JM, Albonetti S, Di Renzo F, Cavani F. Horizons in Sustainable Industrial Chemistry and Catalysis. Vol. 178. Elsevier; 2019. pp. 281-302
  40. 40. Yuan Y-J, Yu Z-T, Chen D-Q, Zou Z-G. Metal-complex chromophores for solar hydrogen generation. Chemical Society Reviews. 2017;46:603-631
  41. 41. Tschierlei S, Karnahl M, Presselt M, Dietzek B, Guthmuller J, González L, et al. Photochemical fate: The first step determines efficiency of H2 formation with a supramolecular photocatalysts. Angewandte Chemie, International Edition. 2010;49:3981-3984
  42. 42. Fukuzumi S, Kobayashi T, Suenobu T. Photocatalytic production of hydrogen by disproportionation of one-electron-reduced rhodium and iridium–ruthenium complexes in water. Angewandte Chemie, International Edition. 2011;50:728-731
  43. 43. Stoll T, Gennari M, Fortage J, Castillo CE, Rebarz M, Sliwa M, et al. An efficient Ru(II) -Rh(III) -Ru(II) polypyridylphotocatalyst for visible-light-driven hydrogen production in aqueous solution. Angewandte Chemie, International Edition. 2014;53:1654-1658
  44. 44. Khine Ma SS, Hisatomi T, Domen K. Hydrogen production by photoctalytic water splitting. Journal of Japan Petroleum Institute. 2013;56(5):280-287
  45. 45. Adams RE, Grusenmeyer TA, Griffith AL, Schmehl RH. Transition metal hydride complexes as mechanistic models for proton reduction catalysis. Coordination Chemistry Reviews. 2018;362:44-53
  46. 46. Berardi S, Drouet S, Franca L, Gimbert-Suriñach C, Guttentag M, Richmond C, et al. Molecular artificial photosynthesis. Chemical Society Reviews. 2014;43:7501-7519
  47. 47. Robinson SJC, Heinekey DM. Hydride & dihydrogen complexes of earth abundant metals: Structure, reactivity, and applications to catalysis. Chemical Communications. 2017;53:669-676
  48. 48. Schilter D, Camara JM, Huynh MT, Hammes-Schiffer S, Rauchfuss TB. Hydrogenase enzymes and their synthetic models: The role of metal hydride. Chemical Reviews. 2016;116:8693-8749
  49. 49. Gärtner F, Sundararaju B, Surkus AE, Boddien A, Loges B, Junge H, et al. Light-driven hydrogen generation: Efficient iron-based water reduction catalysts. Angewandte Chemie, International Edition. 2009;48:9962-9965
  50. 50. Hollmann D, Gärtner F, Ludwig R, Barsch E, Junge H, Blug M, et al. Insights into the mechanism of photocatalytic water reduction by DFT supported in situ EPR/Raman spectroscopy. Angewandte Chemie, International Edition. 2011;50:10246-10250
  51. 51. Gärtner F, Boddien A, Barsch E, Fumino K, Losse S, Junge H, et al. Photocatalytic hydrogen generation from water with iron carbonyl phosphine complexes: Improved water reduction catalysts and mechanistic insights. Chemistry - A European Journal. 2011;17:6425-6436
  52. 52. Gartner F, Cozzula D, Losse S, Boddien A, Anilkumar G, Junge H, et al. Synthesis, characterisation and application of iridium(III) photosensitisers for catalytic water reduction. Chemistry - A European Journal. 2011;17:6998-7006
  53. 53. Yuan YJ, Yu ZT, Gao HL, Zou ZG, Zheng C, Huang W. Tricyclometalated iridium complexes as highly stable photosensitizers for light-induced hydrogen evolution. Chemistry - A European Journal. 2013;19:6340-6349
  54. 54. Lazarides T, Delor M, Sazanovich IV, McCormick TM, Georgakaki I, Charalambidis G, et al. Coutsolelos, Photocatalytic hydrogen production from a noble metal free system based on a water soluble porphyrin derivative and acobaloxime catalyst. Chemical Communications. 2014;50:521-523
  55. 55. Muckerman JT, Fujita E. Theoretical studies of the mechanism of catalytic hydrogen production by a cobaloxime. Chemical Communications. 2011;47(46):12456-12458
  56. 56. Nazeeruddin MK, Diau EWG, Yeh CY, Zakeeruddin SM, Gratzel M. Porphyrin-sensitized solar cells with cobalt (II/III)-based redox electrolyte exceed 12 percent efficiency. Science. 2011;334:629-634
  57. 57. Franco F, Cometto C, Nencini L, Barolo C, Sordello F, Minero C, et al. Local proton source in electrocatalytic CO2 reduction with [Mn(bpy−R)(CO)3Br] complexes. Chemistry - A European Journal. 2017;23:4782-4793
  58. 58. Luca OR, Crabtree RH. Redox active ligands in catalysis. Chemical Society Reviews. 2013;42:1440-1459
  59. 59. Lyaskovskyy V, de Bruin B. Redox non-innocent ligands: Versatile new tools to control catalytic reactions. ACS Catalysis. 2012;22:270-279
  60. 60. Thompson EJ, Berben LA. Electrocatalytic hydrogen production by an aluminium (III)complex: Ligand-based proton and electron transfer. Angewandte Chemie, International Edition. 2015;54:11642-11646
  61. 61. Solis BH, Maher AG, Dogutan DK, Nocera DG, Hammes-Schiffer S. Nickel phlorin intermediate formed by proton-coupled electron transfer in hydrogen evolution mechanism. Proceedings of the National Academy of Sciences of the United States of America. 2015;113:485-492
  62. 62. Luo G-G, Zhang H-L, Tao Y-W, Wu Q-Y, Tian D, Zhang Q. Recent progress in ligand-centered homogeneous electrocatalysts for hydrogen evolution reaction. Inorganic Chemistry Frontiers. 2019;6:343-354
  63. 63. Ran J, Zhang J, Yu J, Jaroniec M, Qiao SZ. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chemical Society Reviews. 2014;43:7787-7812
  64. 64. Berardi S, Drouet S, Francas L, Gimbert-Surinach C, Guttentag M, Richmond C, et al. Molecular artificial photosynthesis. Chemical Society Reviews. 2014;43:7501-7519
  65. 65. Yoshida M, Ueno S, Okano Y, Usui A, Kobayashi A, Kato M. Photochemical hydrogen production from 3d transition-metal complexes bearing o-phenylenediamine ligands. Journal of Photochemistry and Photobiology A: Chemistry. 2015;313:99-106
  66. 66. Konno Y, Matsushita N. pH-dependent color-change behavior of bis(o-phenylenediamine) platinum(II) complex and pH-dependent redox of bis(o-semibenzoquinonediimine)platinum (II) complex. Chemical Society of Japan. 2006;79:1046-1053
  67. 67. Matsumoto T, Chang HC, Wakizaka M, Ueno S, Kobayashi A, Nakayama A, et al. Nonprecious-metal-assisted photochemical hydrogen production from ortho-Phenylenediamine J. American Chemical Society. 2013;135:8646-8654
  68. 68. Roberts GM, Williams CA, Young JD, Ullrich S, Paterson MJ, Stavros VG. Unraveling ultrafast dynamics in photoexcited aniline. Journal of the American Chemical Society. 2012;134:12578-12589
  69. 69. Keffer N, Chavarot-Kerlidou M, Artero V. Hydrogen evolution catalyzed by cobaltdiimine-dioxime complexes Acc. Chemistry Research. 2015;48:1286-1295
  70. 70. Giorgetti M, Berrettoni M, Ascone I, Zamponi S, Seeber R, Marassi R. X-ray absorption spectroscopy study on the electrochemical reduction of Co((DO)(DOH) pn)Br2. Electrochimica Acta. 2000;45:4475-4482
  71. 71. Zhang P, Jacques PA, Chavarot-Kerlidou M, Wang M, Sun L, Fontecave M, et al. Phosphine coordination to a cobalt diimine–dioxime catalyst increases stability during light-driven H2 production. Inorganic Chemistry. 2012;51:2115-2120
  72. 72. Haddad AZ, Garabato BD, Kozlowski PM, Buchanan RM, Grapperhaus CA. Beyond metal-hydrides: Non-transition-metal and metal-free ligand-centered electrocatalytic hydrogen evolution and hydrogen oxidation. Journal of the American Chemical Society. 2016;138:7844-7847
  73. 73. Stubbert BD, Peters JC, Gray HB. Journal of the American Chemical Society. 2011;133:18070-18073
  74. 74. Haddad AZ, Cronin SP, Mashuta MS, Buchanan RM, Grapperhaus CA. Metal-assisted ligand centered electrocatalytic hydrogen evolution upon reduction of a bis(thiosemicarbazonato)Cu(II) complex. Inorganic Chemistry. 2017;56:11254-11265
  75. 75. Zhang P, Wang M, Yang Y, Yao T, Sun L. A molecular copper catalyst for electrochemical water reduction with a large hydrogen-generation rate constant in aqueous solution. Angewandte Chemie, International Edition. 2014;53(50):13806-13807
  76. 76. Wang J, Li C, Zhou Q, Wang W, Hou Y, Zhang B, et al. Photocatalytic hydrogen evolution by Cu(ii) complexes. Dalton Transactions. 2016;45:5439-5443
  77. 77. Das A, Han Z, Brennessel WW, Holland PL, Eisenberg R. Nickel complexes for robust light-driven and electrocatalytic hydrogen production from water. ACS Catalysis. 2015;5:1397-1406

Written By

Shipra Sagar, Ravi K. Kanaparthi, Manish K. Tiwari and Satyen Saha

Submitted: 24 February 2020 Reviewed: 18 May 2020 Published: 23 June 2020