You are on page 1of 60

Photocatalytic water splitting: Quantitative

approaches toward photocatalysis by design

Item Type Article

Authors Takanabe, Kazuhiro

Citation Takanabe K (2017) Photocatalytic water splitting: Quantitative


approaches toward photocatalysis by design. ACS Catalysis.
Available: http://dx.doi.org/10.1021/acscatal.7b02662.

Eprint version Post-print

DOI 10.1021/acscatal.7b02662

Publisher American Chemical Society (ACS)

Journal ACS Catalysis

Rights This document is the Accepted Manuscript version of a Published


Work that appeared in final form in ACS Catalysis, copyright
© American Chemical Society after peer review and technical
editing by the publisher. To access the final edited and published
work see http://pubs.acs.org/doi/10.1021/acscatal.7b02662.

Download date 13/09/2023 16:17:41

Link to Item http://hdl.handle.net/10754/625908


Subscriber access provided by King Abdullah University of Science and Technology Library

Review
Photocatalytic water splitting: Quantitative
approaches toward photocatalysis by design
Kazuhiro Takanabe
ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b02662 • Publication Date (Web): 11 Oct 2017
Downloaded from http://pubs.acs.org on October 18, 2017

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Catalysis is published by the American Chemical Society. 1155 Sixteenth Street
N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 58 ACS Catalysis

1
2
3
4
5
6
7
8
Photocatalytic water splitting: Quantitative
9
10
11
12 approaches toward photocatalysis by design
13
14
15
16
17 Kazuhiro Takanabe*
18
19
20 King Abdullah University of Science and Technology (KAUST), KAUST Catalysis Center
21
22
23
(KCC) and Physical Sciences and Engineering Division (PSE), 4700 KAUST, Thuwal, 23955-
24
25 6900, Saudi Arabia.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
1
ACS Catalysis Page 2 of 58

1
2
3
ABSTRACT. A widely used term, “photocatalysis”, generally addresses photocatalytic
4
5
6 (energetically down-hill) and photosynthetic (energetically up-hill) reactions and refers to the use
7
8 of photonic energy as a driving force for chemical transformations, i.e., electron reorganization
9
10
11
to form/break chemical bonds. Although there are many such important reactions, this
12
13 contribution focuses on the fundamental aspects of photocatalytic water splitting into hydrogen
14
15 and oxygen by using light from the solar spectrum, which is one of the most investigated
16
17
18 photosynthetic reactions. Photocatalytic water splitting using solar energy is considered to be
19
20 artificial photosynthesis that produces a solar fuel because the reaction mimics nature’s
21
22 photosynthesis not only in its redox reaction type but also in its thermodynamics (water splitting:
23
24
25 1.23 eV vs. glucose formation: 1.24 eV). To achieve efficient photocatalytic water splitting, all
26
27 of the parameters, though involved at different timescales and spatial resolutions, should be
28
29
optimized because the overall efficiency is obtained as the multiplication of all these
30
31
32 fundamental efficiencies. The purpose of this review article is to provide the guidelines of a
33
34 concept, “photocatalysis by design”, which is the opposite of “black box screening”; this concept
35
36
37 refers to making quantitative descriptions of the associated physical and chemical properties to
38
39 determine which events/parameters have the most impact on improving the overall
40
41 photocatalytic performance, in contrast to arbitrarily ranking different photocatalyst materials.
42
43
44 First, the properties that can be quantitatively measured or calculated are identified. Second, the
45
46 quantities of these identified properties are determined by performing adequate measurements
47
48 and/or calculations. Third, the obtained values of these properties are integrated into equations so
49
50
51 that the kinetic/energetic bottlenecks of specific properties/processes can be determined, and the
52
53 properties can then be altered to further improve the process. Accumulation of knowledge
54
55
56
ranging in fields from solid-state physics to electrochemistry and the use of a multidisciplinary
57
58
59
60
ACS Paragon Plus Environment
2
Page 3 of 58 ACS Catalysis

1
2
3
approach to conduct measurements and modeling in a quantitative manner are required to fully
4
5
6 understand and improve the efficiency of photocatalysis.
7
8
9
10 KEYWORDS. Photocatalysis, water splitting, electrocatalysis, hydrogen evolution, oxygen
11
12 evolution, band alignment, chemical potential, Fermi level
13
14
15
16
17 Introduction
18
19
20 General strategy for improved photosynthetic reactions: A photocatalyst is a substance that
21
22
23 absorbs photons and generates excited states, which then cause in photophysical and
24
25 photochemical processes as they return to their original ground states.1 Photocatalyst materials
26
27 can consist of additional catalytic components, often called cocatalysts, that catalyze
28
29
30 electrochemical redox reactions.2 Such an electrocatalyst is often essential to the photocatalyst
31
32 (photon absorber) because its surface is not typically designed to catalyze redox reactions unless
33
34 the reaction is an outer-sphere electrochemical reversible reaction. The timescale of
35
36
37 electrocatalysis during the photocatalytic process is sufficiently longer than the timescales of
38
39 photophysical or photochemical processes;3 in many photocatalytic reactions, the photocatalysis
40
41
42
can thus be considered to be electrocatalysis where electrocatalyst components induce the redox
43
44 reactions driven by the potential shifts caused by the photocatalyst (photon absorber).
45
46 Photocatalysis eventually builds an electromotive force (emf), a difference in chemical potentials
47
48
49 or Fermi levels, to enable electrocatalysis.1 This emf transient charging at electrocatalytic sites is
50
51 indeed required for water-splitting photocatalysts because both the hydrogen evolution reaction
52
53 (HER) and oxygen evolution reaction (OER) require multiple electron transfer reactions at the
54
55
56 active species and thus relative slow process compared to prior photophysical processes.
57
58
59
60
ACS Paragon Plus Environment
3
ACS Catalysis Page 4 of 58

1
2
3
One may consider the difference between a photocatalyst (photon absorber and electrocatalyst)
4
5
6 and a device that consists of a photovoltaic and an electrolyzer (PV + E). Making hydrogen by
7
8 PV + E technology is still more expensive than using natural gas reforming.4,5 A large number of
9
10
11
elementary events are common; however, the photocatalyst may induce charge separation
12
13 utilizing an electrocatalyst-semiconductor interface or a solid-liquid junction directly,1
14
15 potentially skipping the p-n junctions in solid-solid structures. This is the major driving force of
16
17
18 cost reduction in photocatalytic system compared to PV + E system. Photocatalytic materials
19
20 also do not require the wiring of a PV (but instead require the collection of produced gases). On
21
22 the other hand, in a PV + E configuration, separating the functions of photovoltaic current
23
24
25 generation (PV) and electrocatalysis (E) is easily optimizable for these two separate components,
26
27 and PV + E systems are therefore expected to produce higher efficiencies than photocatalyst
28
29
systems.6 To combine these systems, PV material can be immersed into aqueous solution,7-9
30
31
32 which allows the material to avoid a detrimental temperature increase (because of the water) and
33
34 the resultant efficiency loss. There is a significant chance, however, that the PV material
35
36
37 corrodes because the water itself is corrosive and even worse at extreme pH values; additionally,
38
39 fewer photons are expected to be absorbed when the PV is in water because, besides photon loss
40
41 due to reflection by the water, absorption coefficient of water is non-zero especially beyond 600
42
43
44 nm.10 Lewis recently reviewed the future possibilities for solar energy conversion technology in
45
46 industry and academia.5 Practical use of either of these systems requires future efforts: PV + E
47
48 prices must be decreased by engineering or some technical advancement that produces a
49
50
51 reasonable device and system, and the photocatalytic efficiency of photocatalysts must be
52
53 improved without forgetting a final, scaled-up reactor design. Recently, immobilization of
54
55
56
photocatalyst powder in sheets for use in overall water splitting was demonstrated,11 but the
57
58
59
60
ACS Paragon Plus Environment
4
Page 5 of 58 ACS Catalysis

1
2
3
overall efficiency of this catalyst remains low, predominantly because of inefficient charge
4
5
6 separation by the photocatalyst materials. Collective efforts are needed to address these complex
7
8 issues in clearly desired solar energy conversion technology.
9
10
11
As recently emphasized by Osterloh,12 photosynthetic reactions (∆G > 0) require detailed
12
13 photon management and charge separation, in contrast to photocatalytic reactions (∆G < 0),
14
15 whose performance is most sensitive to surface area. Basing the selection of materials for photon
16
17
18 absorption (photocatalyst) solely on their bandgap and electrocatalyst, often called cocatalyst, is
19
20 not enough to result in photocatalytic efficiency. We will review that defect density, carrier
21
22 concentrations, and interfaces (metal, semiconductor, electrolyte, etc.) strongly influence
23
24
25 efficiency, even when the same materials are used. This fact is widely known, yet there is no
26
27 consensus as to how to evaluate these properties and consequences. In addition to the use of
28
29
disparate reporting protocols,13-23 this discrepancy is the reason why every research article
30
31
32 reports different efficiencies even when same composition of photocatalyst is used. For instance,
33
34 there are multiple methods to prepare photocatalyst materials,24 but they result in different
35
36
37 photocatalytic performances because unquantifiable or difficult-to-measure properties vary.25
38
39 Photocatalysis research is becoming largely arbitrary because of an infinite number of variables,
40
41 e.g., different precursors, synthesis protocols, annealing, pre/post-treatment, and addition of
42
43
44 small quantities of dopants/impurities/additives (often unconsciously); it is very difficult to
45
46 reproducibly make photocatalysts. In a specific operation, one may want to concretely determine
47
48 how overall efficiency is improved by a particular “quantity”.25 There are excellent review
49
50
51 articles concerning photocatalyst and photoelectrochemical reactions: many focus on various
52
53 materials and techniques of characterization.26-45 This review is specifically targeted to
54
55
56
developing a guideline as to what fundamental key parameters improve photocatalytic
57
58
59
60
ACS Paragon Plus Environment
5
ACS Catalysis Page 6 of 58

1
2
3
efficiencies, regardless of the photocatalyst material. It is time to integrate advanced modeling
4
5
6 into the design of photocatalyst materials. Without these efforts, photocatalytic research remains
7
8 abstract, unestablished and unquantifiable.
9
10
11
Consolidation of chemical potentials and Fermi levels: The basic concept of photocatalysis
12
13 relies on the same protocol as all types of catalysis research: a description of chemical potentials
14
15 of electrons, or Fermi levels. A strong connection between solid-state chemistry and physical
16
17
18 chemistry, or photophysics and electrocatalysis, is the accurate description of chemical potentials
19
20 of electrons in various substances (metals, semiconductors, redox ions in the solution, etc.) at
21
22 thermodynamic equilibrium or under steady-state illumination. The concepts of the chemical
23
24
25 potentials of electrons in metals to semiconductor is well described in an excellent book by
26
27 Sato.46 Each elementary step/event in “catalysis” including photocatalysis, in terms of
28
29
thermodynamics and kinetics, becomes quantitatively describable if we have tools to appraise the
30
31
32 chemical potentials of electrons, especially reactive ones, in molecules, nanoparticles, and solids
33
34 (catalyst materials) and in both reactants and products during (photo)catalysis. Work functions of
35
36
37 metals, Nernstian redox potentials of molecules/ions, and Fermi levels or flatband potentials of
38
39 semiconductors are useful statistical measures of energy equilibrium and flow, although
40
41 overlapping reactive energy states in solids and interfaces makes determining the value of these
42
43
44 potentials difficult. Recent advances in solid-state physics and chemistry establishes reasonable
45
46 theories that can measure (estimate) or calculate such energy levels and their densities of state.
47
48 This estimate, consisting of a large number of quantifiable parameters,25 can ideally be used to
49
50
51 predict overall photocatalytic efficiency, even without experiments. It is thus possible to
52
53 determine which parameter is most influential in improving overall efficiency. This strategy is
54
55
56
57
58
59
60
ACS Paragon Plus Environment
6
Page 7 of 58 ACS Catalysis

1
2
3
one step forward to “photocatalysis by design”; the designing systems to reach a target by
4
5
6 altering specific parameters rather than randomly screening materials.
7
8 Successful photocatalysis requires that charged-up electrocatalysts are maintained at the
9
10
11
potentials where steady-state redox reactions occur. A scheme can be derived, on a scale of the
12
13 chemical potential of electrons (and holes), that visualizes the ideal energy transfer (and loss)
14
15 that occurs during sequential photocatalytic processes. Figure 1 describes an example of the use
16
17
18 of a single semiconductor powder as a photocatalyst that is decorated with HER and OER
19
20 electrocatalysts on the surface, in an attempt to achieve overall water splitting. The process is
21
22 initiated with photon absorption, as depicted in the middle of Figure 1. Upon light absorption, an
23
24
25 excited hole and electron are generated in the valence band and conduction band, respectively,
26
27 on the femtosecond time scale.47 After rapid relaxation to the edges of their respective bands in
28
29
femto- to picoseconds, an exciton (electron-hole pair) is separated into free carriers and the
30
31
32 semiconductor-catalyst interface guides the electron and hole to the HER and OER catalysts,
33
34 respectively, generally in nano- to microseconds.47 Substantial losses of potentials are expected
35
36
37 at the interface (“interfacial loss”) and may originate from entropic contributions of electrons48-50
38
39 and interfacial potential barriers that are generated by inadequate alignment. Successful
40
41 electron/hole transfer to the electrocatalyst shifts the potentials either negatively or positively at
42
43
44 transient time on the millisecond to second time scales, and then maintain steady-state potentials
45
46 that are allowed to drive steady-state electrochemical redox reactions to produce H2 and O2.47
47
48 The solution properties may influence the overall performance by limiting the mass transfer of
49
50
51 the reactant ions. How can we draw this type of scheme for every photocatalyst? What properties
52
53 are involved that determine such potentials at each event? Can we identify the bottleneck that
54
55
56
57
58
59
60
ACS Paragon Plus Environment
7
ACS Catalysis Page 8 of 58

1
2
3
limits the overall efficiency? In the next section, the selected key parameters as well as
4
5
6 photocatalysis events that occur at different time scales are identified.
7
8 List of properties involved in photocatalytic water splitting: The primary effort of this review
9
10
11
focuses on discussing the fundamental parameters that are involved in photocatalytic water
12
13 splitting and their quantitative measurement using powdered semiconductor material (the
14
15 concept can be applied to other photocatalysis as well). Photocatalysis for water splitting indeed
16
17
18 involves a complex series of photophysical and electrocatalytic processes.25 The processes
19
20 involved in photocatalytic reactions are divided into the following six components:
21
22 1. Photon absorption
23
24
25 2. Exciton separation
26
27 3. Carrier diffusion
28
29
4. Carrier transport
30
31
32 5. Catalytic efficiency
33
34 6. Mass transfer of reactants and products
35
36
37 Events 3 and 4 can occur simultaneously and coherently, but are separated here for
38
39 convenience. Figure 2 shows this six-gear concept, which represents the photocatalytic water-
40
41 splitting process sequentially occurring at different time scales.25 Photon absorption initiates non-
42
43
44 equilibrium photophysical and photochemical processes. The photon absorption generates an
45
46 exciton, i.e., excitation of an electron in the valence band (VB) or the highest occupied molecular
47
48 orbital (HOMO) to the conduction band (CB) or the lowest unoccupied molecular orbital
49
50
51 (LUMO).47 The probability to occupy such states are predominantly determined by the electronic
52
53 structure (local displacement of atoms) of the semiconductor. This femtosecond process is
54
55
56
followed by relaxation of the electron and the hole to the bottom of the CB and the top of the
57
58
59
60
ACS Paragon Plus Environment
8
Page 9 of 58 ACS Catalysis

1
2
3
VB, respectively, on a similar time scale.47 Next, the exciton (electron-hole pair) needs to be
4
5
6 separated after overcoming the exciton binding energy determined by the electronic structure;
7
8 this structure should guide the excited electron and hole (polaron) to move independently, being
9
10
11
influenced by their effective masses. The combination of carrier diffusion and transport
12
13 effectively utilizes the introduced interfaces, i.e., potential differences, and successful charge
14
15 transfer typically in microseconds to the electrocatalysts decorated on the surface needs to occur.
16
17
18 Because the kinetics of electrocatalysis are unfortunately sluggish compared to the prior events,
19
20 such electrocatalytically active species will be charged either negatively or positively and drive
21
22 electrocatalytic redox reactions on a time scale typically longer than microseconds.47 The key is
23
24
25 that most of the semiconductor and electrocatalytic properties and measures of efficiency at each
26
27 stage are listed separately and are quantitatively measurable by using various characterization
28
29
and kinetics measurements. Once a material is synthesized, these properties and efficiencies are
30
31
32 quantified so that the bottleneck of the process is identified, leading to improved overall
33
34 efficiency. Previous reports describe the associated equations and measurement protocols in
35
36
37 more detail.25 This contribution aims to emphasize the most influential key components in
38
39 determining overall photocatalytic efficiency.
40
41
42
43
44 Quantification of key properties relevant to photocatalytic water splitting
45
46 Generation rate: When solar energy conversion is the primary concern, analysis of the solar
47
48 spectrum provides useful information regarding theoretical maximum efficiency. The solar-to-
49
50
51 hydrogen (STH) conversion efficiency is defined by the H2-energy generated divided by the
52
53 entire solar irradiance. Using the NREL standard spectrum of AM 1.5G,51 integration of UV
54
55
56
photons accounts for a maximum of 3.3% STH efficiency. Including light from the UV to the
57
58
59
60
ACS Paragon Plus Environment
9
ACS Catalysis Page 10 of 58

1
2
3
visible (to 600 nm) results in a maximum theoretical STH efficiency of 17.8%, while up to 800
4
5
6 nm results in >35% (using a single semiconductor). Analysis of the solar spectrum reveals that
7
8 development of a visible-light-responsive photocatalyst material is essential to achieving
9
10
11
substantial solar energy conversion.52 A representative scheme for various visible light
12
13 responsive materials is shown in Figure 3, which is taken from the review paper by Sivula and
14
15 van de Krol.45 The bandgap of the materials is minimum thermodynamic requirement for high-
16
17
18 efficiency photocatalysis; however, the shape of conduction and valence bands are unique to
19
20 each electronic structure, and densities of state typically become very weak at band edges (bands
21
22 are not rectangular as described below).
23
24
25 It is obvious that if no photons are absorbed, no photocatalysis occurs. The initial step of
26
27 photocatalysis is unambiguously the absorption of a photon and exciton generation by the photon
28
29
absorber. Once the photocatalyst material is chosen for investigation, it is crucial to identify its
30
31
32 electronic structure (displacements of the atoms or a crystal structure), which in turn determines
33
34 the densities of the relevant energy states. Commonly, photon flux of incident light, I0,
35
36
37 commonly lead to the following relationships:
38
39 I 0 = A% + T + RS + S + Rd (1)
40
41
42 where A% is absorptance, the ratio of the absorbed to incident electric field, and T, Rs, S, Rd are
43
44 lights that are transmitted, specularly reflected, forward-scattered and back-scattered,
45
46 respectively.53 Most importantly, the absorption coefficient, α(λ), how far photons of a particular
47
48
49 wavelength can penetrate before it is absorbed by the material, can be measured or calculated as
50
51 a function of wavelength.25,27 It also determines important absorption properties such as the
52
53
54
bandgap, band positions (flatband potentials), and the direct/indirect nature of light absorption.
55
56 The absorption spectra indicate the consequences of bandgap excitation, d-d transitions, phonon
57
58
59
60
ACS Paragon Plus Environment
10
Page 11 of 58 ACS Catalysis

1
2
3
absorptions, and excitations associated with defect states.53 To practically measure absorption
4
5
6 coefficients, the single crystal thin-film configuration of semiconductors provides a more precise
7
8 description because contribution of scattering is minimized and the film thickness is well-
9
10
11
defined.54 From transmittance, T, and reflectance, R,55 values, we obtain α for the film thickness,
12
13 d, when Re−α d << 1 :56
14
15
16
1 T
α = − ln 2
(2)
17 d (1 − R )
18
19
20 The number of electron-hole pairs that are generated per photon striking the semiconductor
21
22 with depth, d, is described using the generation rate, G, per geometric area. After considering
23
24
25 Rayleigh and Mie scattering by the powder, which depend on the size of the particle,57 and
26
27 reflection and transmission by the medium with distinct refractive indices,58 the Beer-Lambert
28
29 law approximation leads to:59
30
31
32 G = α I 0e−α d (3)
33
34
35 From this equation, it is thus obviously essential to obtain photon flux densities by irradiance
36
37 measurements. Other intrinsic parameters can then also be quantified, e.g., the refractive index,
38
39 n; the extinction coefficient, κ(λ); and the dielectric constant, εr, which can be divided into
40
41
42 contributions from the electronic density, ε∞, and from the motion of ions in the material, εvib; εr
43
44 = ε∞ + εvib.61 Methods to obtain these properties can be found in the literature.25
45
46 Once the absorption coefficient is obtained, we obtain the absorption depth, which is a useful
47
48
49 measure how far light can penetrate into a material before being absorbed; the absorption depth
50
51 can be determined by simply taking the inverse of the absorption coefficient α.25 The absorption
52
53
54
depth, together with scattering and reflection, is critical to deciding how thick a photocatalyst
55
56 film or suspension should be or how many semiconductor particles are required to be able to
57
58
59
60
ACS Paragon Plus Environment
11
ACS Catalysis Page 12 of 58

1
2
3
report a useful photocatalytic efficiency. To be able to compare photocatalytic performances
4
5
6 from different laboratories, the maximum photon absorption should be achieved by a
7
8 photoreactor that is used to obtain an “optimal rate” that is not perturbed by the amount of
9
10
11
photocatalyst used.20 Absorption coefficients of typical direct bandgap semiconductors fall into
12
13 the range of 1 × 104 – 1 × 106 cm−1, equivalent to absorption depths of 1000 – 10 nm. A typical
14
15 indirect bandgap semiconductor, Si, possesses a typically low absorption coefficient of 1 × 103 –
16
17
18 1 × 105 cm−1, corresponding to absorption depths of up to a few micrometers for visible light
19
20 (400–800 nm).62 We emphasize that the density of state (DOS) is the primary criterion to select a
21
22 semiconductor for photocatalysis. Essentially, the Franck-Condon principle63,64 suggests that the
23
24
25 displacement of atom positions does not change upon photon absorption and that accurate
26
27 determination of a local crystal structure (Brillouin zone) with an appropriate consideration of
28
29
spin-orbit coupling predominantly determines these optoelectronic properties. Recent advances
30
31
32 in density functional theory (DFT) calculations give quite accurate and reliable estimates of the
33
34 electronic structures and resultant DOS of semiconductors with a given crystal structure.61 When
35
36
37 new photovoltaic and photocatalytic materials are developed, it is recommended that the accurate
38
39 crystal structure, e.g., via Rietveld refinement, which dictates the electronic structure and
40
41 resultant optoelectronic properties, be determined.
42
43
44 As mentioned, single-crystal thin films are preferred in measurements of optoelectronic
45
46 properties because of minimized contribution of scattering and diffuse reflection. It is also noted
47
48 that the Kubelka-Munk function,65 used in diffuse reflectance spectroscopy equipped with an
49
50
51 integrating sphere, is a useful tool for measuring the absorption properties of powder samples.
52
53 However, use of this function often leads to an exaggerated interpretation of the absorption
54
55
56
intensity, especially when close to bandgap. One must consider that near band edges, absorption
57
58
59
60
ACS Paragon Plus Environment
12
Page 13 of 58 ACS Catalysis

1
2
3
coefficients can be exceptionally low, which is not obvious from the spectra plotted using a
4
5
6 Kubelka-Munk function. The absorption spectra and the Kubelka-Munk function also contain
7
8 quantitative information, where a value of zero is especially meaningful. When impurities are
9
10
11
present in the system, non-zero absorption data reflects not only that the spectrum does not
12
13 purely represent the desired compound or material but also the extent of dopant or metallic
14
15 character; therefore, do not forget to plot zero in absorption spectra or Kubelka-Munk function.
16
17
18 The impurity energy levels beyond the bandgap energy also empirically follow the Ulbach rule,66
19
20 which can be additionally considered to quantify the absorption properties of the
21
22 semiconductors.
23
24
25 Exciton binding energy: After successful photon absorption and the resultant exciton
26
27 generation, electron-hole pairs67 are to be separated to generate excited electrons and holes (free
28
29
carriers), or otherwise to recombine easily. The next criterion for selection of a photocatalyst is
30
31
32 the exciton binding energy, which represents the energy required to ionize an exciton from its
33
34 lowest energy state.68 For a Mott-Wannier-type exciton, the 1s state energy, E1, of an exciton
35
36
37 described by the Bohr theory is the exciton binding energy, Rex, and it can be described as
38
39 m∗e4 m∗
40 Rex = = E1 2 (4)
41 2h2ε r2 εr
42
43
44
where εr is a relative permittivity or dielectric constant, m* is the reduced effective mass of the
45
46 1 1 1
47 electron (n)-hole (p) system ( *
= * + * ), e is the elemental charge, and h is Planck’s
m mn mp
48
49
50 constant. A database containing these parameters for a large number of typical semiconductors is
51
52
already available.69 The benchmark energy value is that of thermal energy (25 meV at room
53
54
55 temperature),61 and the efficient separation of excitons requires that the binding energy be lower
56
57 than this value. For Mott-Wannier excitons, the typical binding energy is less than 10 meV, and
58
59
60
ACS Paragon Plus Environment
13
ACS Catalysis Page 14 of 58

1
2
3
the exciton radius is ~10 nm. For Frenkel excitons, such as carbon nitride, these values can be
4
5
6 greater than 1 eV and ~1 nm.70 Such a high exciton binding energy necessitates charge
7
8 separation at the molecular level, similar to the case of bulk heterojunctions of organic
9
10
11
semiconductors. The key properties that affect the values of the exciton binding energy are the
12
13 effective masses and dielectric constant. The effective masses of the electron and hole are
14
15 determined by the curvature of the electronic structure in the conduction and valence bands,
16
17
18 respectively. The electronic dielectric constant is also predominantly determined by the
19
20 electronic structure of a given material. High dielectric materials, such as perovskite structures,
21
22 are typically excellent photocatalysts. Currently, DFT calculation can estimate exciton binding
23
24
25 energies and effective masses as well as different crystal orientations at high accuracy; typically,
26
27 high distortion creates an anisotropic electronic field upon exciton generation, and this field
28
29
assists charge separation.36
30
31
32 Carrier lifetime: For successful photocatalysis, the generated free carriers are transferred to
33
34 redox-active sites on the catalyst surfaces (or to the back contact, in the case of
35
36
37 photoelectrochemistry). The next useful parameter is the minority carrier lifetime, τ, which is
38
39 another intrinsic indicator of whether a semiconductor material can be an effective photocatalyst.
40
41 Generally, the recombination can occur through the following mechanisms; “surface”
42
43
44 recombination, “bulk” Shockley-Read-Hall (defects) (srh),71,72 the band-to-band radiative, and
45
46 the band-to-band Auger (bba),73 and more; the lifetime of which are reciprocally correlated:
47
48
1 1 1 1 1 1 1
49 = + = + + + + ⋅⋅⋅
50 τ τ surf τ bulk τ surf τ srh τ bb τ bba
51
(5)
52
53 Once the carrier lifetime is obtained, recombination rate can be estimated, depending on the
54
55 recombination models. Essentially, the generation rate, which will be described later, will be
56
57
58 canceled out by the recombination rate to leave the effective carrier rates for electrocatalysis.25
59
60
ACS Paragon Plus Environment
14
Page 15 of 58 ACS Catalysis

1
2
3
The carrier lifetime also gives the minority carrier diffusion length, L, representing the average
4
5
6 distance that the excess minority carrier travels from where it was generated to where it is
7
8 annihilated.
9
10
11 L = Dcτ (6)
12
13
14
where Dc is the diffusion coefficient of the carriers, which will be described more in details in
15
16 the following section. The comparison of this value with the particle size of the photocatalyst or
17
18 the film thickness of the photoelectrode is critical in designing the photon absorber.74
19
20
21 An empirical expression between carrier lifetime and dopant concentration was reported by
22
23 Law et al. for indirect bandgap Si (Figure 4), but it describes very interesting trend as a function
24
25 of dopant concentration.75 Two types of srh and bba recombinations are integrated into this
26
27
28 model:
29
30 1 1 1 1 + N D Nref 1
31 = + = + (7)
32 τbulk τ srh τbba τ0 CA N D2
33
34
where τ0 is the low-concentration lifetime, ND is the doping carrier concentration, Nref is the roll-
35
36
37 off concentration, and CA is the Auger coefficient. From this analysis, it is obvious that the
38
39 carrier lifetime increases as the doping level decreases, i.e., the more intrinsic semiconductor
40
41
42 generally has longer carrier lifetimes. This parameter is primarily associated with bulk
43
44 recombination that originates from the srh process and impurity concentrations (and therefore,
45
46 strictly speaking, the descriptor is not the carrier concentration alone). GaAs semiconductor
47
48
49 shows similar trend.76 Relatedly, the surface treatment (etching or shell formation) of single-
50
51 crystal surfaces substantially improves the minority carrier lifetime, as measured by time-
52
53 resolved photoluminescence, or the photoconductivity lifetime, as measured by terahertz
54
55
56
57
58
59
60
ACS Paragon Plus Environment
15
ACS Catalysis Page 16 of 58

1
2
3
photoconductivity. Examples of this behavior include that of well-investigated photovoltaic
4
5
6 semiconductors such as CdTe,77,78 InP,79 and GaAs.80
7
8 In the case of powder samples or nano-structured architectures, whose surface contributions
9
10
11
are large relative to that of the bulk, the surface states largely influence the minority carrier
12
13 lifetime. The carrier concentration is obviously an important factor in photocatalysis, but what is
14
15 the relevant quantity in the semiconductor powders? When a highly defective material with a
16
17
18 carrier concentration of 1018 cm−3 is used, a typical 10-nm cube of photocatalyst contains only
19
20 one carrier per photocatalyst particle in the bulk. If fewer dopant materials with a carrier
21
22 concentration of 1016 cm−3 can be synthesized, a 50-nm cube of photocatalyst contains one
23
24
25 carrier per particle, and a 100-nm cube photocatalyst contains ten carriers. Since the contribution
26
27 from the bulk is so low, contributions from the surface are easily observed.42 The surfaces and
28
29
interfaces are, by definition, defects, often possessing non-stoichiometric and/or dangling bonds
30
31
32 that are electronically active. The potential of the surface states is a difficult-to-measure quantity:
33
34 surface dipoles associated with functional groups, adsorbate, and redox ions in electrochemical
35
36
37 double layer all affect. The consequence is, however, known that they create intermittent states
38
39 within bandgap or HER-OER redox potentials.81 Therefore, the surface not only acts as a charge
40
41 separation interface but also as a recombination site of excited carriers such as impurities and
42
43
44 abrupt terminations. What is the concentration of such surface sites? In the case of an oxide
45
46 material, the maximum concentration should be comparable to the surface hydroxyl site density,
47
48 which typically reaches ~4 nm−2, depending on the identity of the oxide, its treatment, and
49
50
51 characteristics of the exposed facets.82 On a fully hydroxylated surface, there are ~60,000 surface
52
53 sites on a 50-nm cube-shaped particle, giving a surface site to bulk carrier ratio >100. From this
54
55
56
number, it is clear that management of surface states is critical to controlling the photocatalytic
57
58
59
60
ACS Paragon Plus Environment
16
Page 17 of 58 ACS Catalysis

1
2
3
activity of powder materials.83 If surface sites are insulated or deactivated to the point at which
4
5
6 their concentration is 100 times less than that of complete exposure, the surface and bulk atoms
7
8 of the same system become comparable. This idea is visualized in Figure 5, and it is consistent
9
10
11
with the common observation that having “high crystallinity” and a minimal number of defects
12
13 enhances photocatalytic efficiency, in contrast to a simple increase in surface area.84 In summary,
14
15 the minority carrier lifetime is prolonged with a more intrinsic semiconductor (fewer dopants) in
16
17
18 the bulk, and surface modification at the interface is crucial for photocatalysts, which is further
19
20 interacted with the following parameters.
21
22 Carrier diffusion and transport: The important parameters to consider when selecting
23
24
25 photocatalyst materials are, at this point, predominantly the electronic structure, which
26
27 determines the absorption coefficient, and the charge carrier concentration, which influences
28
29
carrier lifetime and diffusion length. The next event that occurs during the photocatalytic process
30
31
32 is excited carrier transport. Charge separation is a primary concern for photosynthetic reaction
33
34 and solid liquid interface should be effectively utilized.12 The generated free charge carriers must
35
36
37 travel through the bulk of semiconductor to the surface redox sites.85 Such phenomena can be
38
39 described in terms of electron flow, i.e., current. There are two driving forces for electron (n) and
40
41 hole (p) movement: diffusion driven by concentration gradient and drift driven by potential
42
43
44 gradient:67
45
46 J = J diffusion + J drift (8)
47
48
49 J n = eDn ∇ n + ne µ n E (9)
50
51
J p = −eDp∇p + peµp E (10)
52
53
54 where e is the elementary charge, D is the diffusion coefficient, ∇ p and ∇ n are the gradients of
55
56
57 electrons or holes, µ is the mobility of the charge carrier, p denotes hole concentration, n denotes
58
59
60
ACS Paragon Plus Environment
17
ACS Catalysis Page 18 of 58

1
2
3
electron concentration, and E is electric field. The diffusion contribution is associated with the
4
5
6 diffusion coefficient and mobility of intrinsic semiconductors:61
7
8 k BT
9 D= µ (11)
10 e
11
12 where kB is the Boltzmann constant. The mobility in a specific direction can be further described
13
14 by
15
16
17 τc
18
µ=e (12)
m*
19
20
21 where τc is the collision time of the charge carrier and m* is the effective mass. The criterion for
22
23 good mobility under ambient condition is considered to be m* < 0.5me (e for electron).61 As
24
25 mentioned previously, the electronic structure predominantly determines the effective masses
26
27
28 and thus the mobility and diffusion coefficient. Practically, the resistivity, charge carrier
29
30 concentration and resultant mobility of the semiconductors can be measured by using the van der
31
32
33
Pauw technique with Hall measurements,86,87 although this method is better when using a high-
34
35 quality semiconductor slab. If there is no potential gradient, free carriers are transferred via
36
37 diffusion, which is a very inefficient form of carrier transport. The minority carrier diffusion
38
39
40 length can be as short as a few nm, but only when the carrier lifetime is on the order of
41
42 picoseconds.
43
44 Therefore, movement of free charge carriers must be adequately guided by potential gradients,
45
46
47 generating drift current. Such gradients can be made by effective utilization of metal-
48
49 semiconductor, semiconductor-electrolyte, and semiconductor-semiconductor interfaces
50
51
including surface modifications.85 The decoration of the surfaces of semiconductors causes
52
53
54 several effects: the reduction of surface recombination; the introduction of potential gradient; and
55
56 modification with catalytic components.
57
58
59
60
ACS Paragon Plus Environment
18
Page 19 of 58 ACS Catalysis

1
2
3
At the metal-semiconductor interface, the key parameters that control the energy level are the
4
5
6 work function of the metal and the Fermi level of the semiconductor,46 which may result in a
7
8 Schottky barrier or ohmic contact, depending on their relative positions and the carrier
9
10
11
concentrations. For details, please refer to, e.g., the work of Tung.88 In the literature,89 barrier
12
13 heights at semiconductor-metal interface were correlated with electronegativity of metal and
14
15 nature of semiconductor, either ionic (Si, Ge, etc.) or covalent (oxides, like TiO2, SrTiO3, etc.).
16
17
18 Figure 6 shows representative interesting trends of barrier heights, ϕBn0, for various metals and
19
20 semiconductors. An index, “S”, the slope of Figure 6A, gives sensitivity of electronegativity of
21
22 metal, XM, to the barrier heights. Relatively small S for ionic semiconductors shows that barrier
23
24
25 heights are insensitive to electronegativity (or workfunction) of metals, whereas large S for
26
27 covalent semiconductors indicate they are more sensitive to the difference between Fermi level
28
29
of semiconductor and metal workfunction. Ohmic contacts have been reported for various
30
31
32 combinations of semiconductor and metals and metal alloys.89 Such smooth contacts may
33
34 improve efficiency by bridging excited electrons to electrocatalysts. Practically, Ti is commonly
35
36
37 used as a contact layer for p-Si,91 and Ti and Ta are used for some covalent materials, such as
38
39 SrTiO392 and LaTiO2N.93 For an unique case, the Cu2O photocathode achieved high
40
41 photocathodic current when successive deposition of ZnO:Al and TiO2 before Pt catalyst
42
43
44 deposition.94 Powder semiconductor seem to be more challenging to achieve this type of
45
46 decoration in nanoscale, so the establishing technique that allow to develop the smooth contact
47
48 may lead to high efficiency.
49
50
51 At the semiconductor-electrolyte interface,95,96 the Fermi level of the semiconductor and the
52
53 reduction potential of the solution play a crucial role in determining potential gradient. A
54
55
56
successful application of the solid-electrolyte interfaces is the dye-sensitized solar cell, where
57
58
59
60
ACS Paragon Plus Environment
19
ACS Catalysis Page 20 of 58

1
2
3
TiO2 collects excited electrons to its conduction band from the dyes anchored to its surface.97 A
4
5
6 key to avoid charge recombination is a band bending of TiO2, guiding the injected electrons to its
7
8 back contact.74 Similarly, photocatalyst surface will experience the band bending when
9
10
11
immersed in water. Solving the Poisson equation to x-direction (eq. 13) leads to description of
12
13 band bending, and this space charge layer should be utilized to achieve effective charge
14
15 separation.67
16
17
18 d 2Φ x eN
19 2
=− D (13)
dx ε 0ε r
20
21
22 where Φx is the potential as a function of x, ND is the majority carrier density, ε0 is the static
23
24 permittivity in vacuum, εr is the static relative permittivity or dielectric constant of the
25
26
27 semiconductor. The key parameters in determining the space charge layer are the carrier
28
29 concentration and the dielectric constant of the semiconductor. The electrolyte is strongly
30
31 influenced by the surface state and potential-determining ions at the surface.98 In water, the
32
33
34 isoelectric point of the semiconductor provides a useful indication of whether the surface is
35
36 negatively or positively charged.99 Semiconductor-semiconductor interfaces can form p-n
37
38
junctions, but the details regarding this process are described elsewhere.67
39
40
41 The consequence of potential gradients at interfaces account for the photovoltage: the origin of
42
43 emf of the electrocatalysts, determining primary efficiency of the photocatalytic system. It is
44
45
46 emphasized that the bandgap is not equivalent with the photovoltage; substantial potential losses
47
48 are expected at surfaces and other interfaces. The Si bandgap of 1.1 eV typically gives an open-
49
50 circuit voltage or photovoltage gain of only 0.7 eV (~40 % loss) in photovoltaic system.100
51
52
53 Therefore, reporting the bandgap is not likely to be sufficient in further understanding
54
55 photocatalytic processes and material properties. At the same time, band alignment of various
56
57 materials is a good start to discuss: however, the Fermi level equilibration between two materials
58
59
60
ACS Paragon Plus Environment
20
Page 21 of 58 ACS Catalysis

1
2
3
(p-n or n-n junctions with type I and II alignments, etc.) does not result in smooth interface,
4
5
6 which is strongly influenced by lattice match, impurities, degree of atom diffusion in mutual
7
8 phases at interface (related to annealing). At this moment, there seems only empirical choices to
9
10
11
achieve least-barrier interface for most cases, but in the case of simplified bulk semiconductor,
12
13 there is a theory to predict the electronic structure, which certainly helps the guideline for
14
15 material design.101-106
16
17
18 Simplified two-dimensional numerical modeling, a widely known calculation in solar cell
19
20 community, is able to describe potential gradients inside the semiconductor using classical
21
22 semiconductor device equations.39,107 These simulations can provide reasonable estimates of
23
24
25 quantum efficiency and STH efficiency as a function of wavelength. The quantification of
26
27 several parameters, i.e., absorption coefficient, band positions, dielectric constant, carrier
28
29
concentrations, effective masses, mobility, and lifetime, has been discussed thus far. The beauty
30
31
32 of this modeling is that the sensitivity of the fundamental parameters can be investigated and the
33
34 properties that are most influential in determining the overall photocatalytic efficiencies can be
35
36
37 identified, i.e., modeling brings us one step closer to “photocatalysis by design”. The
38
39 overpotentials that are required for HER and OER on the surface are input variables (future work
40
41 is required to make them outputs of the modeling) in this approach,107 and the diffusion-drift
42
43
44 current equation can be solved using generation and recombination rates when the system is
45
46 under steady-state illumination. In this way, the influence of the metal dispersion on the
47
48 photocatalytic performance can also be evaluated.99 For many of the equations that are involved
49
50
51 in the estimation of photovoltaic currents, the readers are referred to previous studies.25,107
52
53 A scheme in Figure 7 shows how the potential gradient close to HER electrocatalyst on an n-
54
55
56
type semiconductor under steady-state illumination may look in the bulk of the semiconductor
57
58
59
60
ACS Paragon Plus Environment
21
ACS Catalysis Page 22 of 58

1
2
3
when carrier concentrations, carrier lifetimes and carrier mobilities are varied.107 The
4
5
6 semiconductor surface was decorated with HER catalyst particle that collects excited electrons,
7
8 assuming ohmic contact at the interface. The photocatalyst surface was designed to oxidize
9
10
11
water, where Schottky contact with electrolyte was assumed. On the photocatalyst surface (left
12
13 side of Figure 7A), it was assumed that substantial potential gradient exists between HER
14
15 catalyst and semiconductor bare surface to achieve overall water splitting (1.53 eV). In Figure
16
17
18 7B, excited electrons should flow from right (semiconductor bulk) to left (surface) and
19
20 downwards, following the slope generated at the semiconductor-electrolyte interface. At high
21
22 carrier concentrations, a substantial energy barrier (peak), related to the so-called pinch-off
23
24
25 effect,88 was observed close to the surface, even if ohmic contact was assumed. On the other
26
27 hand, lower carrier concentrations result in gradual slopes that guide the excited electron to the
28
29
left side of the HER catalyst. This observation coincides with finding resultant larger AQE and
30
31
32 STH efficiencies at lower carrier concentrations. This type of modeling certainly helps
33
34 determining the properties that should be targeted to improve overall performance; e.g., the
35
36
37 carrier lifetime should be greater than hundreds of picoseconds, etc.
38
39 Similar simulations suggest that using defective materials, such as dispersions (in particle size
40
41 and density) of metal nanoparticles, on semiconductor absorbers does not significantly influence
42
43
44 efficiency, although a metal catalyst is essential in achieving effective charge separation.107 This
45
46 result suggests that even though photocatalytic efficiency remains the same, the turnover
47
48 frequency (TOF; e.g., rate per surface metal site) varies with different electrocatalyst
49
50
51 dispersions.23 This variation is because, in this case, charge separation efficiency determines
52
53 overall efficiency, and electrocatalysts only consume carriers as they arrive. The potential,
54
55
56
determined as a consequence of charge separation, is different for each particle, and the current
57
58
59
60
ACS Paragon Plus Environment
22
Page 23 of 58 ACS Catalysis

1
2
3
(rate) per particle is also thus different; i.e., photocatalysts are electrocatalysts, so the rate
4
5
6 (current) is based on the potential.23 A smaller particle, or greater surface area, does not
7
8 necessarily result in the best overall STH efficiency, and moreover, a high TOF does not always
9
10
11
lead to high photocatalytic efficiency of the entire system. On contrary, it is expected that high
12
13 exciton binding energy materials may only require high dispersion of catalyst to create more
14
15 number/density of interfaces. As a result, the modeling provides guidance whether photocatalyst
16
17
18 properties should be altered or the identity or dispersion of the electrocatalyst should be
19
20 improved, which is another step forward to “photocatalysis by design”.
21
22 As seen above, a quantitative description of such optoelectronic properties can be used to
23
24
25 estimate theoretical photocatalytic efficiency in ideal semiconductor situations. It gives, at
26
27 minimum, a good estimate whether the improvement of a semiconductor (including its interface)
28
29
or an electrocatalyst should be investigated and even which specific parameters should be
30
31
32 altered, such as minority carrier concentrations or the catalyst dispersion on the surface.23 It also
33
34 may allow researchers to consider the potential loss associated at the interfaces. There are
35
36
37 measurement techniques that can be used to estimate the potential drop at the (oxy)hydroxide
38
39 layer on semiconductor surfaces.108-112 It is already effective to simply isolate bare photocatalyst
40
41 surface from the water electrolyte, e.g., by using some oxide (e.g., SiO2, Al2O3, or TiO2), thus
42
43
44 avoiding the surface state and the photocorrosion that is prevalent in some semiconductor
45
46 compounds.91,94,113-115 However, precisely describing the potential at the interface is still under
47
48 development. For example, the classic model fails to describe realistic porous ion-permeable
49
50
51 electrocatalysts (i.e., oxyhydroxide cocatalysts for water oxidation). Classical semiconductor
52
53 equations are applicable to bulk materials (the smallest particle size in the COMSOL model used
54
55
56
above was 100 nm).107 The smaller particles have a higher specific surface area, a shorter travel
57
58
59
60
ACS Paragon Plus Environment
23
ACS Catalysis Page 24 of 58

1
2
3
distance to the surface for their charge carriers, a lower degree of band bending, and, possibly, a
4
5
6 wider band gap because of quantum size effects.83 Substantial efforts on efficient interfacial
7
8 development have been made, especially in photoelectrochemistry applications, and one may
9
10
11
refer to a recent excellent review by Li and co-workers on this topic.116 Nevertheless, diffusion
12
13 and drift remain fundamental principles that describe carrier transport from the bulk to the
14
15 surface. Simple simulations, as described above, already predict a substantial loss in the potential
16
17
18 gradient at the surface. Successfully incorporating anisotropy in electronic structures of crystals
19
20 by a simple manner (simpler than conventional fabrication of a p-n junction) is the way to make
21
22 photocatalysts more efficient and cheaper.83 Other directions may include preventing electrolyte
23
24
25 junctions from inducing surface recombination as well as improving the majority carrier
26
27 pathway, e.g., by using a metal-insulator-semiconductor-type junction or a carefully embedded
28
29
buried-junction active-site that is electronically isolated from environmental effects.117,118
30
31
32 Paradoxically, an approach to insulating a semiconductor from solution is by actually
33
34 minimizing the beneficial possible utilization of band bending at solid-liquid interfaces. Unlike
35
36
37 photoelectrochemical measurement, photocatalysis using powder photon absorber cannot apply
38
39 external electric field: i.e., the electrons and holes must find their own way to lead to
40
41 electrocatalysts. A breakthrough to boost the photoconversion efficiency resides in unique
42
43
44 establishment of the interface bridging photon absorber and electrocatalyst.
45
46 Electrocatalytic activity: The climax of the water-splitting reaction finishes with the successful
47
48 consumption of the photogenerated charge carriers by electrochemical redox reactions. The
49
50
51 mismatch between the time scales of charge transfer and electrocatalysis causes accumulation of
52
53 electrons/holes at their respective redox-active species at the surface, resulting in potential shifts
54
55
56
(transient charge up) on metal or metal (oxy)hydroxide particles or, in some cases, the
57
58
59
60
ACS Paragon Plus Environment
24
Page 25 of 58 ACS Catalysis

1
2
3
photocatalyst surface itself, as a catalyst component.REF Measurements of such potential shifts
4
5
6 were reported for metal particles by using probe molecules under illumination.119-123 and for
7
8 metal (hydr)oxides using electrochemical techniques.109-112 At given potentials, the catalysts
9
10
11
should electrocatalyze HER and OER, respectively; the performances of these reactions can be
12
13 separately measured by using electrochemical techniques that can determine the exact values of
14
15 applied potentials.1 Electrocatalysis is another unique interface event:124 The reaction proceeds
16
17
18 on the catalyst surface atoms together with electrolyte within a double-layer region where at least
19
20 three-water-equivalent ions/molecules are involved in covalent and non-covalent nature. This
21
22 requires consideration of not only inner Helmholtz layer but outer Helmholtz layer to describe,
23
24
25 e.g., transition states, which means that counter “supporting ions” play significant role in
26
27 electrocatalytic kinetics.125 It is tremendously difficult, if not possible, to precisely describe the
28
29
chemical potentials at double-layer region,126 but various efforts are ongoing as electrocatalysis
30
31
32 is indeed a core technology to convert renewable energy resources to useful chemical forms.5
33
34 Electrocatalytic water splitting itself is a field of study in which many efforts are currently
35
36
37 ongoing. Electrocatalytic activity can be ranked using the quantitative values of the exchange
38
39 current of a given catalyst, i0, and the transfer coefficient, α, which are described by the Tafel
40
41 equation when the reverse reaction is neglected (Eq. 11);1 however, the terms of these extracted
42
43
44 values are still ambiguous due to the lack of a method that can precisely determine active surface
45
46 areas.
47
48
49 i0 α nF ( Ecat − E 0 )
50 r= exp , (14)
nF RT
51
52
53 where n is the number of electrons involved in the reaction, F is the Faraday constant, Ecat and E0
54
55 are the Fermi level of the catalyst and the redox potential in solution, respectively, R is the
56
57
58
universal gas constant, and T is the absolute temperature. The overpotential is defined as the
59
60
ACS Paragon Plus Environment
25
ACS Catalysis Page 26 of 58

1
2
3
difference between Ecat and E0, an additional voltage required (relative to the thermodynamic
4
5
6 potential) to drive the respective redox reaction. Microkinetic Tafel analysis for water redox
7
8 chemistries has been reviewed elsewhere.127 Key aspects of the fundamental study of HERs and
9
10
11
OERs include finding descriptors of electrocatalytic activity. Based on Sabatier’s principle,
12
13 metal-hydrogen bond strengths characterize the HER exchange current density.127,128 Generally,
14
15 the OER catalyst is also characterized by using the metal-oxygen bond strength as a descriptor,
16
17
18 because a linear relationship exists among metal-oxygen, metal-hydroxide and metal-
19
20 oxyhydroxide bond strengths, all of which may be involved in the rate-determining steps.129-131
21
22 When acidic conditions are chosen, the development of a non-noble metal electrodes with acid
23
24
25 tolerance is required. The recent development of metal phosphide materials is of significant
26
27 interest because they contain only abundant transition metals, such as Ni, Fe, and Co.132-134 For
28
29
OERs, mixed oxyhydroxides,135 such as nickel-iron,136,137 perovskites,138 and spinels139 have also
30
31
32 been reported as low overpotential electrocatalysts in alkaline conditions that do not use noble
33
34 metals. One must remember that for industrial applications, catalyst durability is often more
35
36
37 important than catalytic performance.9 “Self-healing” capability, i.e., dissolution-redeposit
38
39 process of the electrocatalyst during electro- and photocatalysis is a compelling method to
40
41 achieve long-term durability.140-144 Additionally, the temperature of the solution in a practical
42
43
44 photoreactor may be substantially higher than room temperature because the photoreactor may
45
46 absorb infrared irradiation; this factor should be considered in experiments regarding activation
47
48 energy. Ironically, high activation energies lead to a highly sensitive current increase with
49
50
51 temperature changes, resulting in excellent performance under certain relevant water-splitting
52
53 conditions (e.g., 10 mA cm−2 at mild pH for Ni vs. NiFe OER catalysts)144 or enthalpy-entropy
54
55
56
compensations (e.g., in the case of HER at mild pH for Pt, Ni, NiCu, etc.).145
57
58
59
60
ACS Paragon Plus Environment
26
Page 27 of 58 ACS Catalysis

1
2
3
One of the unique features of photocatalysis is that the photon absorber materials are not stable
4
5
6 under extreme pH conditions (acidic or alkaline), which is the regime in which commercial
7
8 electrolyzers are operated.146 Interestingly, pure water without any supporting electrolyte can be
9
10
11
used in overall water splitting with a powder semiconductor photocatalyst147,148 because of the
12
13 very short distance between the HER (cathode) and the OER (anode), which should occur on the
14
15 same surface with minimum solution resistance. Electrochemistry is a powerful tool to
16
17
18 quantitatively evaluate reactions under near-neutral (or mild) pH conditions.146 The impact of pH
19
20 on these reactions will be discussed in greater detail later. In short, one must first identify the
21
22 “reactants” of respective redox reactions: pH change will cause “reactant switching” at a given
23
24
25 current level, which is associated with the diffusion contribution.149 In general, hydronium ions
26
27 (protons) are more easily reduced than water molecules,149 and hydroxyl ions are more readily
28
29
oxidized than water molecules.146 One critical note is that the “water-splitting” reaction rather
30
31
32 paradoxically does not prefer water molecules to be its reactant.146,149 This preference results
33
34 from the fact that water molecules contain very strong O-H bonds (as is also obvious from the
35
36
37 fact that H2O is one of the most thermodynamically stable compounds). To facilitate the water
38
39 molecule dissociation, anisotropic sites on the surface are effective for heterolytic dissociation of
40
41 water molecules. For example, in alkaline conditions, Markovic and co-workers reported that
42
43
44 islands of nickel or cobalt species on noble metal surfaces (such as Pt) further enhance both
45
46 HERs and OERs.150
47
48 Knowledge obtained by electrocatalytic studies should be successfully transferred to the
49
50
51 photocatalytic studies. In any case, electrocatalysis should catch up with current flow from the
52
53 electrons and holes that are generated in the semiconductor underneath. It is obvious that small
54
55
56
potential shifts should trigger the corresponding current flow, and excellent electrocatalysts are
57
58
59
60
ACS Paragon Plus Environment
27
ACS Catalysis Page 28 of 58

1
2
3
thus preferred as efficient photocatalysis.52 An excellent example is rough CoOx modification on
4
5
6 n-Si photoanode, where almost full utilization of the generated photovoltage was achieved for
7
8 OER by CoOx under illumination.151 If there is difference in “ranking” electrocatalyst materials
9
10
11
during photocatalysis and pure electrocatalysis, it arises from different degrees of potential shifts
12
13 at the catalyst/semiconductor interfaces, causing the electrocatalysts to not experience the same
14
15 potentials due to, e.g., different degree of Schottky contact and barrier height, as discussed
16
17
18 above, Figure 6.52 On contrary approach, if the kinetics of electrocatalysts for HER and OER
19
20 require substantial overpotentials (e.g., overvoltage of ~2.0 V), the required bandgap to
21
22 maximize theoretical STH efficiency essentially becomes larger; the best scenario as large as
23
24
25 ~2.4 eV, thus never reaching 10% STH benchmarking efficiency.153 Excellent activity of
26
27 electrocatalysts that are optically transparent is desired.
28
29
To achieve efficient overall water splitting in a membrane-less configuration, as Gerischer
30
31
32 stated in his early work,85 the suppression of the back reactions of H2 and O2 to form H2O must
33
34 be suppressed. Noble metals, in particular, are generally excellent HER catalysts but also
35
36
37 typically catalyze the back reaction either thermally or electrocatalytically (oxygen reduction
38
39 reaction).154,155 Successful suppression with nanometer-scale decorations on such electrocatalyst
40
41 surfaces (core@shell structure) have been reported and use chromium,156,157 molybdenum,158
42
43
44 titanium,159 and lanthanoids,160 as shells. The amorphous structure of very small hydrated
45
46 clusters makes the materials function as a selective membrane that is not permeable to dissolved
47
48 gases (including O2), thus preventing back reactions.156,158 There is a possibility to utilize this
49
50
51 functionality to protect the surface from poisoning because this membrane function also insulates
52
53 various redox-active species.157 Development of shell materials that make the OER catalyst
54
55
56
selective is also ongoing.
57
58
59
60
ACS Paragon Plus Environment
28
Page 29 of 58 ACS Catalysis

1
2
3
It is interesting to note that overall photocatalytic efficiency may be further improved by
4
5
6 having better HER or OER electrocatalysts under all conditions. Based on this charge-up theory,
7
8 enhancements in the rate of reduction or oxidation improve the overall efficiency of water
9
10
11
splitting, which is determined by the photon flux and the efficiency of carrier transport from the
12
13 photocatalyst to the redox catalysts because accelerated electron or hole processes affect the
14
15 potential, which in turn perturbs the rates of the process on the opposite side.161 Because electron
16
17
18 and hole transport are parallel reactions, the overall photocatalysis process does not have a single
19
20 rate-determining step unlike the case of half-reaction electrocatalysis).44,161 In other words,
21
22 further improvement for electron consumption (HER) or hole consumption (OER) should
23
24
25 improve overall efficiencies. Fast consumption of electrons will cause hole accumulation on the
26
27 OER side, further enhancing the overall rate, or vice versa.162,163 It is also effective to analyze the
28
29
sensitivity of HER or OER performance to overall photocatalysis performance, which can be
30
31
32 evaluated using photocatalysis with isotope effects44 or effectively comparing electrocatalytic
33
34 reactions under dark conditions.164
35
36
37 Mass transfer (ion diffusion): After decades of studying photocatalysis for water splitting, the
38
39 efficiency of this process has been tremendously improved.165 The primary focus of
40
41 photosynthetic reactions is still based on managing photons in the bulk and on the surfaces of
42
43
44 photon absorber materials, as mentioned previously. The research in this field has therefore been
45
46 largely oriented by the synthesis of efficient photon absorber materials, including their
47
48 electrocatalyst decorations. Nevertheless, when reaching commercially viable efficiency is
49
50
51 considered, the mass transfer of reactants and ions in addition to solution resistance can no
52
53 longer be ignored during electrocatalysis or photocatalysis.150 Much research is focused on
54
55
56
systems that work at room temperature, and under such conditions, it is often the case that the
57
58
59
60
ACS Paragon Plus Environment
29
ACS Catalysis Page 30 of 58

1
2
3
diffusion of ions contributes to the overall efficiency by creating a concentration overpotential;
4
5
6 an additional loss originates from the depletion of reactants. This is certainly the case when
7
8 photocatalytic remediation of low-concentration substances is the target reaction.12 The rigorous
9
10
11
and quantitative determination of parameters in such a process is essentially possible using the
12
13 thermodynamic and kinetic information that can be generated using electrochemistry.
14
15 At a given current, sources of potentials are classified into kinetic overpotential (dependent on
16
17
18 catalyst), concentration overpotential (independent of catalyst), and solution resistance. The
19
20 contributions of the concentration overpotential and solution resistance can be quantitatively
21
22 obtained by using the physical properties of the solution. Detailed quantification and
23
24
25 methodology was reviewed in previous literature. Mass transport phenomena in electrochemistry
26
27 are described by the Nernst-Planck equation with terms for (in this order) diffusion, migration
28
29
and convection (for species i in the x-direction):1
30
31
32 ∂ai ( x ) zi F ∂φ ( x )
33 J i ( x ) = − Di − Di ai − ai v ( x )
34
∂x RT ∂x (15)
35
36 where J is the flux, D is the diffusion coefficient, z is the charge number and v is the velocity of
37
38
the forces in the solution.1 The Stokes-Einstein model gives the diffusion coefficient as
39
40
41 kT
D=
42 3π d µ
43 (16)
44
45 where k is the Boltzmann constant, d is the effective diameter of the ion in the hydrated form
46
47 (Stokes diameter) and µ is the viscosity of the solution.166 Therefore, the parameters governing
48
49
50 mass transport flux are the effective size of the species, viscosity of the solution, and activity (or
51
52 fugacity) of the species. Moreover, solubility of dissolved gases, another quantifiable parameter,
53
54
55
is greatly influenced by the identity and molarity of the supporting electrolyte, which correlates
56
57 with reverse reaction of products going back to water, or hydrogen oxidation reaction (HOR) and
58
59
60
ACS Paragon Plus Environment
30
Page 31 of 58 ACS Catalysis

1
2
3
oxygen reduction reaction (ORR).167 These physical properties can be obtained separately, often
4
5
6 from a database,168 and the contribution of mass transport is thus quantifiable. Under relevant
7
8 reaction conditions, the benchmark STH efficiency of 10% corresponds to a hydrogen
9
10
11
production rate of ~154 µmol H2 cm−2 h−1 and a corresponding current of ~8.3 mA cm−2
12
13 (assuming that a single semiconductor (or tandem semiconductors) is achieving the overall water
14
15 splitting). Under static conditions (no convection) at 25 °C, even hydronium (proton) and
16
17
18 hydroxide ions can face diffusion-limiting currents, causing “reactant switching”: pH values of
19
20 ~1.6 or lower (for hydronium ion) and ~12.3 or higher (for hydroxide ion) are necessary in
21
22 unbuffered conditions. Outside this range of pH (unbuffered, near-neutral pH), reactant
23
24
25 switching between H+ (HER) and OH− (OER) to H2O must occur, causing additional kinetic
26
27 overpotential.150 Obviously, this activity of the reactants, together with minimized solution
28
29
resistance, is one reason why extreme pH conditions are chosen for the electrolysis of water. In
30
31
32 addition, the HER causes an increase in the pH, and the OER causes a decrease in the pH, so the
33
34 complete isolation of ions in a two-compartment cell will lead to a high concentration
35
36
37 overpotential (shifting the thermodynamic potential by 59 mV pH−1), which causes additional
38
39 loss of overall efficiency. The use of an ion-exchangeable membrane is mandatory to separate H2
40
41 and O2 while minimizing these concentration overpotentials.169 Nafion or an alkaline membrane
42
43
44 typically works in media with extreme pH,170 although some membranes that may be used as
45
46 neutral pH values have been recently developed.171 One of the most significant benefits of co-
47
48 producing an H2/O2 mixture is avoiding the use of membranes and minimizing solution
49
50
51 resistance and the pH gradient, i.e., the concentration overpotential.169 However, this process
52
53 occurs at the expense of producing an explosive gas mixture (H2 and O2).169
54
55
56
57
58
59
60
ACS Paragon Plus Environment
31
ACS Catalysis Page 32 of 58

1
2
3
Near-neutral pH conditions makes it possible to use many materials for stable photocatalytic
4
5
6 water splitting. Under neutral pH conditions, buffering ions are commonly used to maintain local
7
8 pH values by utilizing their buffering action.172 In addition to the role of the supporting
9
10
11
electrolyte in minimizing solution resistance (iR drop), the buffer ions significantly influence
12
13 electrocatalysis. The buffer’s counter anion is a carrier for H+ and thus plays a role in
14
15 transporting the H+ reactant to cathode or abstracting the H+ products from the anode.173-175 At
16
17
18 relevant current densities and under ambient conditions, the diffusion of buffer ions may
19
20 therefore result in substantial concentration overpotentials. For example, using an excellent
21
22 catalyst, such as Pt, for HER in NaH2PO4, the optimum buffer concentrations for the HER appear
23
24
25 to be as high as 1.5-2.0 M at 25 °C.174 It has also been recently determined that some OER
26
27 catalysts also suffer from concentration overpotentials, or mass-transfer limitation of buffering
28
29
ions, which can be seen because the rotating-disk electrode current depends on its rotation
30
31
32 rate.174 This fact is often neglected in the photocatalysis and photoelectrochemistry community.
33
34 Another interesting consideration of the activity of the reactants is the use of water vapor as a
35
36
37 reactant (liquid water vs. water vapor). Using vapor-phase water has advantages such as the
38
39 ability to easily control its supply and the use of simple reactor designs, e.g., a fixed bed for
40
41 powder systems.175 However, using water vapor as a reactant encounters considerable difficulties
42
43
44 due to an additional term for the adsorption of water vapor at low partial pressures, which may
45
46 strongly decrease the overall efficiency. In contrast, using liquid-phase water (close to unity) or
47
48 the associated ions (H+ or OH−) as reactants can result in high activities, effectively utilizing the
49
50
51 electric field applied at the double-layer region.
52
53 In static photoelectrochemical water splitting, high efficiency (currents) of photoelectrodes
54
55
56
leads to additional efficiency loss due to the generated gas bubbles blocking the surface. It was
57
58
59
60
ACS Paragon Plus Environment
32
Page 33 of 58 ACS Catalysis

1
2
3
reported that hydrophilic surface is preferred to detach the generated bubbles, which was also
4
5
6 confirmed by the fact that introduction of surfactant was thus effective to remove gas bubbles.
7
8 Contact angle at gas-surface interface is known to be correlated, but the details can be found
9
10
11
elsewhere. For photocatalyst powder systems, the gas bubble problems seem absent in the
12
13 literature, yet probably because of low efficiency of the powder suspension system. Since the
14
15 photocatalytic performance is being improved, one must consider the final form of photocatalyst
16
17
18 samples, whether they should be immobilized in which substrates, and types of convective flow
19
20 of liquids. Associated with this, the temperature of the reactant water may be considered, which
21
22 impact not only catalytic rates but also the mass transport which also has activation energy with
23
24
25 Arrhenius relation.
26
27 Discussion and perspectives
28
29
This contribution identified a number of quantifiable parameters associated with the complex
30
31
32 processes that occur during photocatalytic water splitting. The processes are sequentially and
33
34 often coherently connected in the following order and operate at different time scales as the
35
36
37 following six-gear concept as shown in Figure 2. These parameters and relevant useful
38
39 information are summarized in Table 1. Strategy should be adequately planned to investigate
40
41 photocatalytic materials and reactions. In photoelectrochemical water splitting study, there are
42
43
44 some suggested guidelines in the literature,176 which are also useful for investigating
45
46 photocatalytic water splitting as many phenomena function in common principles. Most of the
47
48 constants and the quantifiable variables are listed in Table 2. By identifying these “quantities”,
49
50
51 one may predict efficiencies using various established equations and thus help directing the
52
53 researches. For further details regarding the equations, the readers are referred to previous
54
55
56
reviews for semiconductor study,25,107,177 and electrochemistry study,173 and literature cited
57
58
59
60
ACS Paragon Plus Environment
33
ACS Catalysis Page 34 of 58

1
2
3
therein. One may want to check the absorption coefficient (Gear 1), especially close to band
4
5
6 edges, the exciton binding energy (Gear 2) and carrier lifetime (Gear 3). Next is to describe
7
8 carrier diffusion (Gear 3) and transport (Gear 4). In simple cases, simulation can currently
9
10
11
estimate maximum photocatalytic quantum efficiencies based on the quantified values of bulk
12
13 semiconductor parameters. This will be the first assessment whether the bulk semiconductor
14
15 properties should be ever suitable for efficient photocatalysis. The guidance obtained is whether
16
17
18 the material itself should be altered, the synthesis protocol should be improved (crystallinity), or
19
20 electrocatalyst decoration should be improved in terms of dispersion and loading, etc.
21
22 In a typical example, Si is widely available already in a commercial scale, and, upon
23
24
25 purchasing, many parameters mentioned in the first part (Gears 1-4) are effectively quantified. Si
26
27 wafers are commercially available with known dopant concentration and conductivity/resistivity.
28
29
A µm-order diffusion length is accordingly obtainable, so the strategy is to focus on the further
30
31
32 improvement in optical enhancement and surface charge separation with dispersed catalyst
33
34 decoration. A deconvolution of such properties (including Gear 4-6) for 3D structure p-Si
35
36
37 photocathode was well reported by Esposito and co-workers.178 On contrary, for another
38
39 example, non-oxide materials, such as Ta3N5 used as a visible-light-responsive photocatalyst,
40
41 have many parameters unknown and remain uncertainties to the semiconductor properties. Such
42
43
44 quantities of the parameters were summarized in the literature for Ta3N5.179 This type of
45
46 (oxy)nitride materials is synthesized at each laboratory usually via nitridation of oxide precursors
47
48 in NH3 flow at high temperatures. Non-stoichiometry due to remaining oxygen as well as anion
49
50
51 vacancy associated with Ta5+ reduction to Ta3+ in the bulk structure is also recognized.180 From
52
53 the literature, despite the efforts for the surface alteration and catalyst decoration essential to
54
55
56
enhance photocatalytic performance, there remain a lot to do to improve bulk properties and
57
58
59
60
ACS Paragon Plus Environment
34
Page 35 of 58 ACS Catalysis

1
2
3
improve carrier lifetimes.107 Diffusion lengths can be as short as a few nm when carrier trapping
4
5
6 of Ta3N5 happens in a picosecond-order181,182 (although a part of carriers that have longer
7
8 lifetimes was also reported for Ta3N5).183,184 It is still important that the synthesis protocol should
9
10
11
be improved for prolong carrier lifetime and resultant diffusion lengths; e.g., flux-assisted
12
13 protocols for improved crystallinity.
14
15 It is also time to thoroughly discuss the actual reaction conditions with practical reactor design
16
17
18 for photocatalysis (pressure, temperature, activity, etc.),185-189 because they will affect the
19
20 photovoltage, electrocatalytic kinetics, and the diffusion of ions, etc. As a result, the
21
22 performance/durability ranking of photocatalytic/electrocatalytic materials may be different from
23
24
25 investigations at room temperature under low pressures. Accordingly, the era has come for solar
26
27 fuel production study to seriously consider the practical photoreactor design together with
28
29
reaction conditions.186 At a photoreactor in large scale application under solar irradiation,
30
31
32 temperature may substantially rise by design, which may be beneficial because many
33
34 photocatalytic systems are reported to have positive activation energy,11 and kinetic isotope
35
36
37 effects from D2O experiments;189,190 i.e., surface electrocatalysis may be sluggish enough to
38
39 influence overall efficiency. It is desired to maintain small density and high dispersion of the
40
41 catalysts on the photon absorber not to absorb photons by the catalysts themselves. Such small
42
43
44 quantities of the catalysts, especially when the efficiency is improved, prefers high temperature
45
46 to catch up the electrocatalysis. Choice of materials should also be conducted based on durability
47
48 and robustness of the photocatalyst materials under the operational reaction conditions.
49
50
51 This contribution may serve as a set of guidelines to help identify the kinetic bottleneck with
52
53 “quantities” that limit the overall efficiency of photocatalysis and to help intentionally improve
54
55
56
specific properties: steps forward toward “photocatalysis by design” concept. Finally, the
57
58
59
60
ACS Paragon Plus Environment
35
ACS Catalysis Page 36 of 58

1
2
3
concepts shown here are not limited to this reaction and may be applied to, e.g., photocatalytic
4
5
6 CO2 reduction or even environmental remediation. To jump to a commercial level of
7
8 photocatalytic efficiency, consolidated efforts to achieve commercial solar energy conversion
9
10
11
processes based on an understanding at the microscopic and macroscopic levels should be made.
12
13
14
15 AUTHOR INFORMATION
16
17
18 Corresponding Author
19
20
* Kazuhiro Takanabe, email: kazuhiro.takanabe@kaust.edu.sa
21
22
23
24 ACKNOWLEDGMENT
25
26 The research reported in this publication was supported by King Abdullah University of Science
27
28
29 and Technology (KAUST). The author appreciates Dr. Angel T. Garcia-Esparza for thorough
30
31 discussion on simulation data related to Figure 7.
32
33
34 REFERENCES
35
36 (1) Electrochemical methods, 2nd Ed.; Bard, A. J.; Faulkner L. R., Eds.; John Wiley & Sons,
37
38
Inc.: Danvers, 2001; pp 736-768.
39
40 (2) Kraeutler, B.; Bard, A. J. J. Am. Chem. Soc. 1978, 100, 2239-2240.
41
42 (3) Calvo, E. J. In Electrode kinetics: principles and methodology, Bamford, C. H.; Tipper, C. F.
43 H.; Compton R. G., Eds.; Elsevier: Amsterdam, 1986; Vol. 26, pp. 1-74.
44
45 (4) Shaner, M. R.; Atwater, H. A.; Lewis, N. S.; McFarland, E. W. Energy Environ. Sci. 2016, 9,
46 2354-2371.
47
48 (5) Lewis, N. S. Science 2016, 351, aad1920.
49
50 (6) Nakamura, A.; Ota, Y.; Koike, K.; Hidaka, Y.; Nishioka, K.; Sugiyama, M.; Fujii, K Appl.
51 Phys. Express 2015, 8, 107101.
52
53 (7) Khaselev, O.; Turner, J. A. Science, 1998, 280, 425-427.
54
55 (8) Nocera, D. G. Acc. Chem. Res. 2012, 45, 767-776.
56
57
58
59
60
ACS Paragon Plus Environment
36
Page 37 of 58 ACS Catalysis

1
2
3
(9) Sun, K.; Saadi, F. H., Lichterman, M. F.; Hale, W. G.; Wang, H.-P.; Zhou, X.; Plymale, N.
4
5 T.; Omelchenko, S. T.; He, J.-H; Papadantonakis, K. M.; Brunschwig, B. S.; Lewis, N. S. Proc.
6 Nat. Am. Sci. 2015, 112, 3612-3617.
7
8 (10) Kageshima, Y.; Shinagawa, T.; Kuwata, T.; Nakata, J.; Minegishi, T.; Takanabe, K.;
9 Domen, K. Sci. Rep. 2016, 6, 24633.
10
11 (11) Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z; Takata, T.;
12 Nakabayashi, M.; Shibata, N.; Li, Y.; Sharp, I. D.; Kudo, A.; Yamada, T.; Domen, K. Nat.
13
Mater. 2016, 15, 611-615.
14
15
16
(12) Osterloh, F. E. ACS Energy Lett. 2017, 2, 445-453.
17
18
(13) Mills, A.; Wang, J. J. Photochem. Photobio. A 1999, 127, 123-134.
19
20
(14) Yang, X.; Ohno, T.; Nishijima, K.; Abe, R.; Ohtani, B. Chem. Phys. Lett. 2006, 429, 606-
21 610.
22
23 (15) Ohtani, B. Chem. Lett. 2008, 37, 217-229.
24
25 (16) Mills, A Appl. Catal. B 2012, 128, 144-149.
26
27 (17) Kisch, H. Angew. Chem. Int. Ed. 2013, 52, 812-847.
28
29 (18) Schanze, K. S.; Kamat, P. V.; Buriak, J. M. ACS Appl. Mater. Interfaces. 2014, 6, 11815-
30 11816.
31
32 (19) Buriak, J. M. Chem. Mater. 2014, 26, 2211-2213.
33
34 (20) Bahnemann, D.; Kisch, H. J. Phys. Chem. Lett. 2015, 6, 1907-1910.
35
36 (21) Coridan, R. H.; Nielander A. C.; Francis, S. A; McDowell, M. T.; Dix, V.; Chatman, S. M;
37 Lewis, N. S. Energy Environ. Sci. 2015, 8, 2886-2901.
38
39
(22) Buriak, J. M.; Jones, C. W.; Kamat, P. V.; Schanze, K. S.; Schatz, G. C.; Scholes, G. D.;
40
41 Weiss, P. S. Chem. Mater., 2016, 28, 3525-3526.
42
43 (23) Qureshi, M.; Takanabe, K. Chem. Mater. 2017, 29, 158-167.
44
45 (24) Takanabe, K.; Domen, K. ChemCatChem 2012, 4, 1485-1497.
46
47 (25) Takanabe, K. Top. Curr. Chem. 2016, 371, 73-103.
48
49 (26) Nozik, A. J. Ann. Rev. Phys. Chem. 1978, 29, 189-222.
50
51 (27) Nosaka, Y.; Ishizuka, Y.; Miyama, H. Ber. Bunsenges. Phys. Chem. 1986, 90, 1199-1204.
52
53 (28) Memming, R. Top. Curr. Chem. 1988, 143, 79-112.
54
55 (29) Hagfeldt, A.; Grätzel, M. Chem. Rev. 1995, 95, 49-68.
56
57
58
59
60
ACS Paragon Plus Environment
37
ACS Catalysis Page 38 of 58

1
2
3
(30) Photocatalysis science and technology, Kaneko, M; Okura, I., Eds.; Kodansha-Springer:
4
5 Tokyo-New York, 2002: pp 1-360.
6
7 (31) Domen, K. In Photocatalysis – Heterogeneous. Encyclopedia of Catalysis, Horvath, I. T.
8 Ed., Wiley, Weinheim, 2002.
9
10 (32) Maeda, K.; Domen, K. J. Phys. Chem. C 2007, 111, 7851-7861.
11
12 (33) Kamat, P. V. J. Phys. Chem. C 2007, 111, 2834-2860.
13
14 (34) Osterloh, F. E. Chem. Mater. 2008, 20, 35-54.
15
16 (35) Kudo, A.; Miseki, Y. Chem. Soc. Rev. 2009, 38, 253-278.
17
18 (36) Inoue, Y. Energy Environ. Sci. 2009, 2, 364-386.
19
20 (37) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. A.;
21 Lewis, N. S. Chem. Rev. 2010, 110, 6446-6473.
22
23 (38) Maeda, K.; Domen, K. J. Phys. Chem. Lett. 2010, 1, 2655-2661.
24
25 (39) Hisatomi, T.; Minegishi, T.; Domen, K. Bull. Chem. Soc. Jpn. 2012, 85, 647-655.
26
27 (40) Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Adv. Mater. 2012, 24, 229-
28
29 251.
30
31 (41) Tachibana, Y.; Vayssieres, L.; Durrant, J. R. Nat. Photonics 2012, 6, 511-518.
32
33 (42) Osterloh, F. E. Chem. Soc. Rev. 2013, 42, 2294-2320.
34
35 (43) Hisatomi, T.; Kubota, J.; Domen, K. Chem. Soc. Rev. 2014, 43, 7520-7535.
36
37 (44) Hisatomi, T.; Takanabe, K.; Domen, K. Catal. Lett. 2014, 145, 95-108.
38
39 (45) Sivula, K.; van de Krol, R. Nat. Rev. Mater. 2016, 1, 15010.
40
41 (46) Sato, N. Electrochemistry at metal and semiconductor electrodes, Sato, N. Eds; Elsevier,
42 Amsterdam, 1998: pp 1-396.
43
44 (47) Modern molecular photochemistry of organic molecules, Turro, N. J.; Ramamurthy, V.;
45 Scaiano, J. C. Eds.; University Science Books, Sausalito, 2010; pp 1-38.
46
47 (47) Polman, A.; Atwater, H. A. Nat. Mater. 2012, 11, 174-177.
48
49 (48) Osterloh, F. E. J. Phys. Chem. Lett. 2014, 5, 3354-3359.
50
51 (49) Omelchenko, S. T.; Tolstova, Y.; Atwater, H. A.; Lewis, N. S. ACS Energy Lett. 2017, 2,
52
53
431-437.
54
55 (51) Absorption and scattering of light by small particles, Bohren, C. F.; Huffman, D. R. Eds.;
56 Wiley-VCH, Weinheim, 2004: pp 1-476.
57
58
59
60
ACS Paragon Plus Environment
38
Page 39 of 58 ACS Catalysis

1
2
3
(52) Interpreting diffuse reflectance and transmittance, Dahm, D. J.; Dahm, K. D. Eds.; NIR
4
5 publications: West Sussex, 2007: pp 1-286.
6
7 (53) Bae, D.; Pedersen, T.; Seger, B.; Malizia, M.; Kuznetsov, A.; Hansen, O.; Chorkendorff, I.;
8 Vesborg, P. C. K. Energy. Environ. Sci. 2015, 8, 650-660.
9
10 (55) National Renewable Energy Laboratory (NREL) website:
11 http://rredc.nrel.gov/solar/spectra/am1.5
12
13 (56) Takanabe, K.; Domen, K. Green 2011, 1, 313-322.
14
15 (57) Chen. Z.; Dinh, H. N.; Miller, E. Photoelectrochemical Water Splitting Standards,
16 Experimental Methods, and Protocols. Springer, New York. 2013.
17
18 (58) Wemple, S. H.; Seman, J. A. Appl. Opt. 1973, 12, 2947-2949.
19
20 (59) Di Giulio, M.; Micocci, G.; Rella, R.; Siciliano, P.; Tepore, A Phys. Stat. Sol. a 1993, 136.
21
22 K101-K104.
23
24 (60) Lodenquai, J. F. Solar Energy 1994, 53, 209-210.
25
26 (61) Le Bahers, T.; Rérat, M.; Sautet, P. J. Phys. Chem. C 2014, 118, 5997-6008.
27
28 (62) Green, M. A. Solar Energy Mater Solar Cells 2008, 92, 1305-1310.
29
30 (63) Franck, J. Trans. Farad. Soc. 1926, 21, 536-542.
31
32 (64) Condon, E. Phys. Rev. 1926, 28, 1182-1201.
33
34 (65) Džimbeg-Malčić, V.; Barbarić-Mikočević, Ž.; Itrić, K. Technical Gazette 2011, 18, 117.
35
36 (66) Kurik, M. V. Phys. Stat. Sol. a 1971, 8, 9-45.
37
38 (67) Introduction to Solid State Physics, 8th Ed.; Kittel, C. Ed.; Wiley, Weinheim; 2005: pp 1-
39 704.
40
41 (68) Bastard, G.; Mendez, E. E.; Chang, L. L.; Esaki, L. Phys. Rev. B 1982, 26, 1974-1979.
42
43 (69) Semiconductors: Data Handbook, 3rd ed.; Madelung, O.; Springer: New York, 2004.
44
45 (70) Melissen, S.; Le Bahers, T.; Steinmann, S. N.; Sautet, P. J. Phys. Chem. C 2014, 119,
46 25188-25196.
47
48
(71) Shockley, W.; Read, W. T. Jr. Phys. Rev. 1952, 87, 835-842.
49
50
51
(72) Hall, R. N. Phys. Rev. 1952, 87, 387.
52
53
(73) Auger, P. C. R. A. S. 1952, 177, 169.
54
55 (74) Leng, W. H.; Barnes, P. R. F.; Juozapavicius, M.; O’Regan, B. C.; Durrant, J. R. J. Phys
56 Chem. Lett. 2010, 1, 967-972.
57
58 (75) Law, M. E.; Solley, E.; Liang, M.; Burk, D. E. IEEE Electron Dev. Lett. 1991, 12, 401-403.
59
60
ACS Paragon Plus Environment
39
ACS Catalysis Page 40 of 58

1
2
3
(76) Ito, H.; Furuta, T.; Ishibashi, T. Appl. Phys. Lett. 1991, 58, 2936-2938.
4
5
6
(77) Cohen, R.; Lyahovitskaya, V Appl. Phys. Lett. 1998, 73, 1400-1402.
7
8
(78) Kuciauskas, D.; Kanevce, A.; Dippo, P.; Seyedmohammadi, S.; Malik, R. IEEE J.
9 Photovoltaics, 2015, 5, 366-371.
10
11 (79) Rosenwaks, Y.; Shapira, Y, Phys. Rev. B 1992, 45, 9108-9119.
12
13 (80) Casey, H. C. Jr.; Miller, B. I.; Pinkas, E. J. Appl. Phys., 1973, 44, 1281-1287.
14
15 (81) Lewis, N. S. Inorg. Chem. 2005, 44, 6900-6911.
16
17 (82) Zhuravlev, L. T. Coll. Surf. A 2000, 173, 1-38.
18
19 (83) Yoneyama, H. Crit. Rev. Sold State Mater. Sci. 1993, 18, 69-111.
20
21 (84) Fukasawa, Y.; Takanabe, K.; Shimojima, A.; Antonietti, M.; Domen, K.; Okubo, T. Chem.
22 Asian J. 2011, 6, 103-109.
23
24 (85) Gerischer, H. J. Phys. Chem. 1984, 88, 6096-6097.
25
26 (86) van der Pauw, L. J. Philips Res. Rep. 1958, 13, 1.
27
28 (87) Heaney, M. B. Electrical conductivity and resistivity. In: The measurement, instrumentation
29
and sensors handbook. CRC, Boca Raton, 2000.
30
31
32
(88) Tung, R. T. Appl. Phys. Rev. 2014, 1, 011304.
33
34
(89) Physics of semiconductor devices, 3rd Ed., Sze, S. M.; Ng, K. K. Eds.; Wiley: New York,
35 2006: pp 134-196.
36
37 (90) Kurtin, S.; McGill, T. C.; Mead, C. A. Phys. Rev. Lett. 1969, 22, 1433-1436.
38
39 (91) Esposito, D. V.; Levin, I.; Moffat, T. P.; Talin, A. A. Nat. Mater. 2013, 12, 562-568.
40
41 (92) Ham, Y.; Minegishi, T.; Hisatomi, T.; Domen, K. Chem. Commun. 2016, 52, 5011-5014.
42
43 (93) Minegishi, T.; Nishimura, N.; Kubota, J.; Domen, K. Chem. Sci. 2013, 4, 1120-1124.
44
45 (94) Paracchino, A.; Laporte, V.; Sivula, K.; Grätzel, M.; Thimsen, E. Nat. Mater. 2011, 10, 456-
46 461.
47
48 (95) Gerischer, H. Electrochim. Acta 1990, 35, 1677-1699.
49
50 (96) Zhang, Z.; Yates, J. T. Jr. Chem. Rev. 2012, 112, 5520-5551.
51
52 (97) Grätzel, M. Nature 2001, 414, 338-344.
53
54 (98) Chamousis, R. L.; Osterloh, F. E. Energy Environ. Sci. 2014, 7, 736-743.
55
56 (99) Butler, M. A.; Ginley, D. S. J. Electrochem. Soc. 1978, 125, 228-232.
57
58
59
60
ACS Paragon Plus Environment
40
Page 41 of 58 ACS Catalysis

1
2
3
(100) Martin A. Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E. W.; Levi, D.
4
5 H.; Ho-Baillie, A. W. Y. Prog. Photovolt. Res. Appl. 2017, 25, 668-676.
6
7 (101) Turner, J. A. J. Chem. Educ. 1983, 60, 327-329.
8
9 (102) Lewis, N. S. J. Electrochem. Soc. 1984, 131, 2496-2503.
10
11 (103) Rossi, R. C.; Tan, M. X.; Lewis, N. S. Appl. Phys. Lett. 2000, 77, 2698-2700.
12
13 (104) Rossi, R. C.; Lewis, N. S. J. Phys. Chem. B 2001, 105, 12303-12318.
14
15 (105) Bisquert, J.; Cendula, P.; Bertoluzzi, L.; Gimenez, S. J. Phys. Chem. Lett. 2014, 5, 205-
16 207.
17
18 (106) Xiang, C.; Weber, A. Z.; Ardo, S.; Berger, A.; Chen, Y.; Coridan, R.; Fountaine, K. T.;
19 Haussener S.; Hu, S.; Liu R.; Lewis, N. S.; Modestino, M. A.; Shaner, M. M.; Singh, M. R.;
20
Stevens, J. C.; Sun K.; Walczak, K. Angew. Chem. Int. Ed. 2016, 55, 12947-12988.
21
22
(107) Garcia-Esparza, A. T.; Takanabe, K. J. Mater. Chem. A 2016, 4, 2894-2908.
23
24
25
(108) Cendula, P.; Tilley, S. D.; Gimenez, S.; Bisquert, J.; Schmid, M.; Grätzel, M.;
26 Schumacher, J. O. J. Phys. Chem. C 2014, 118, 29599-29607.
27
28 (109) Mills, T. J.; Lin, F.; Boettcher, S. W. Phys. Rev. Lett. 2014, 112, 148304.
29
30 (110) Lin, F.; Boettcher, S. W. Nat. Mater. 2014, 13, 81-86.
31
32 (111) Lichterman, M. F.; Hu, S.; Richter, M. H.; Crumlin, E. J.; Axnanda, S.; Favaro, M.;
33 Drisdell, W.; Hussain, Z.; Mayer, T.; Brunschwig, B. S.; Lewis, N. S.; Liu, Z.; Lewerenz, H. J.
34 Energy Environ. Sci. 2015, 8, 2409-2416.
35
36 (112) Laskowski, F. A. L.; Nellist, M. R.; Venkatkarthick, R.; Boettcher, S. W. Energy Environ.
37 Sci. 2017, 10, 570-579.
38
39
(113) Hu, S.; Shaner, M. R.; Beardslee, J. A.; Lichterman, M.; Brunschwig, B. S.; Lewis, N. S.
40
41 Science 2014, 344, 1005-1009.
42
43 (114) Verlage, E.; Hu, S.; Liu, R.; Jones, R. J. R.; Sun, K.; Xiang, C.; Lewis, N. S.; Atwater, H.
44 A. Energy Environ. Sci. 2015, 8, 3166-3172.
45
46 (115) Zhou, X.; Liu, R.; Sun, K.; Chen, Y.; Verlage, E.; Francis, S. A.; Lewis, N. S.; Xiang, C.
47 ACS Energy Lett. 2016, 1, 764-770.
48
49 (116) Ding, C.; Shi, J.; Wang, Z; Li, C ACS Catal. 2017, 7, 675-688.
50
51 (117) Hill, J. C.; Landers, A. T.; Switzer, J. A. Nat. Mater. 2015, 14, 1150-1156.
52
53 (118) Digdaya, I. A.; Adhyaksa, G.; Trzesniewski, B. J.; Garnett, E.; Smith, W. A. Nat.
54 Commun. 2017, 8, 15968.
55
56 (119) Kamat, P. V. Pure Appl. Chem. 2002, 74, 1693-1706.
57
58
59
60
ACS Paragon Plus Environment
41
ACS Catalysis Page 42 of 58

1
2
3
(120) Jakob, M.; Levanon, H.; Kamat, P. V. Nano. Lett. 2003, 3, 353-358.
4
5
6
(121) Subramanian, V.; Wolf, E. E.; Kamat, P. V. J. Am. Chem. Soc. 2004, 126, 4943-4950.
7
8
(122) Yoshida, M.; Yamakata, A.; Takanabe, K.; Kubota, J.; Osawa, M.; Domen, K. J. Am.
9 Chem. Soc. 2009, 131, 13218-13219.
10
11 (123) Lu, X.; Bandara, A.; Katayama, M.; Yamakata, A.; Kubota, J.; Domen, K. J. Phys. Chem.
12 C 2011, 115, 23902-23907.
13
14 (124) Stamenkovic, V. R.; Strmcnik, D.; Lopes, P. P.; Markovic, N. M. Nat. Mater. 2017, 16,
15 57-69.
16
17 (125) Suntivich, J.; Perry, E. E.; Gasteiger, H. A.; Shao-Horn, Y. Electrocatal. 2013, 4, 49-55.
18
19 (126) Fuel cell catalysis, a surface science approach, Koper, M. T. M., Ed.; Wiley, Hoboken;
20
2009: pp 1-158.
21
22
(127) Shinagawa, T.; Garcia-Esparza, A. T.; Takanabe, K. Sci. Rep. 2015, 5, 13801.
23
24
25
(128) Trasatti, S. J. Electroanal. Chem. 1972, 32, 163-184.
26
27
(129) Greeley, J.; Jaramillo, T. F.; Bonde, J.; Chorkendorff, I.; Nørskov, J. K. Nat. Mater. 2006,
28 5, 909-913.
29
30 (130) Matsumoto, Y.; Sato, E. Mater. Chem. Phys. 1986, 14, 397-426.
31
32 (131) Man, I. C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H. A.; Martinez, J. I.; Inoglu, N. G.;
33 Kitchin, J.; Jaramillo, T. F.; Nørskov, J. K.; Rossmeisl, J. ChemCatChem 2011, 3, 1159-1165.
34
35 (132) Grimaud, A.; May, K. J.; Carlton, C. E.; Lee, Y. L.; Risch, M.; Hong, W. T.; Zhou, J.;
36 Shao-Horn, Y. Nat. Commun. 2013, 4, 3439.
37
38 (133) Popczun, E. J.; McKone, J. R.; Read, C. G.; Biacchi, A. J.; Wiltrout, A. M.; Lewis, N. S.;
39
Schaak, R. E. J. Am. Chem. Soc. 2013, 135, 9267-9270.
40
41
(134) Jiang, P.; Liu, Q.; Liang, Y.; Tian, J.; Asiri, A. M.; Sun, X. Angew. Chem. Int. Ed. 2014,
42
43 53, 12855-12859.
44
45 (135) Popczun, E. J.; Read, C. G.; Roske, C. W.; Lewis, N. S.; Schaak, R. E. Angew. Chem. Int.
46 Ed. 2014, 53, 5427-5430.
47
48 (136) Smith, R. D. L.; Prévot, M. S.; Fagan, R. D.; Zhang, Z.; Sedach, P. A.; Siu, J. M. K.;
49 Trudel, S.; Berlinguette, C. P. Science 2013, 340, 60-63.
50
51 (137) Gong, M.; Li, Y.; Wang, H.; Liang, Y.; Wu, J. Z.; Zhou, J.; Wang, J.; Regier, T.; Wei, F.;
52 Dai, H. J. Am. Chem. Soc. 2013, 135, 8452-8455.
53
54 (138) Gong, M.; Zhou, W.; Tsai, M. C.; Zhou, J.; Guan, M.; Lin, M. C.; Zhang, B.; Hu, Y.;
55
56
Wang, D. Y.; Yang, J.; Pennycook, S. J.; Hwang, B. J.; Dai, H. Nat. Commun. 2014, 5, 5695.
57
58
59
60
ACS Paragon Plus Environment
42
Page 43 of 58 ACS Catalysis

1
2
3
(139) Suntivich, J.; May, K. J.; Gasteiger, H. A.; Goodenough, J. B.; Shao-Horn, Y. Science
4
5 2011, 334, 1383-1385.
6
7 (140) Wei, C.; Feng, Z.; Scherer, G. G.; Barber, J.; Shao-Horn, Y.; Xu, Z. J. Adv. Mater. 2017,
8 29, 1606800.
9
10 (141) Kanan, M. W.; Nocera, D. G. Science 2008, 321, 1072-1075.
11
12 (142) Lutterman, D. A.; Surendranath, Y.; Nocera, D. G. J. Am. Chem. Soc. 2009, 131, 3838-
13 3839.
14
15 (143) Dinca, M.; Surendranath, Y.; Nocera, D. G. Proc. Natl. Acad. Sci. 2010, 107, 10337-
16 10341.
17
18 (144) Kuang, Y.; Jia, Q.; Ma, G.; Hisatomi, T.; Minegishi, T.; Nishiyama, H.; Nakabayashi, M.;
19
20 Shiata, N.; Yamada, T.; Kudo, A.; Domen, K Nat. Energy, 2016, 2, 16191.
21
22 (145) Shinagawa, T.; Ng, M. T.-K.; Takanabe, K. Angew. Chem. Int. Ed. 2017, 56, 5061-5065.
23
24 (146) Shinagawa, T.; Takanabe, K. J. Phys. Chem. C 2016, 120, 24187-24196.
25
26 (147) Electrochemistry 2nd Ed, Hamann, C. H.; Hamnett, A.; Vielstich, W. Eds.; Wiley-VCH,
27 Weinheim, 2007; pp. 397-438.
28
29 (148) Kato, H.; Asakura, K.; Kudo, A. J. Am. Chem. Soc. 2003, 125, 3082-3089.
30
31 (149) Sakata, Y.; Matsuda, Y.; Nakagawa, T.; Yasunaga, R.; Imamura, H.; Teramura, K.
32 ChemSusChem 2011, 4, 181-184.
33
34 (150) Shinagawa, T.; Garcia-Esparza, A. T.; Takanabe, K. ChemElectroChem 2014, 1, 1497-
35 1507.
36
37 (151) Subbaraman, R.; Tripkovic, D.; Chang, K. C.; Strmcnik, D.; Paulikas, A. P.; Hirunsit, P.;
38
39 Chan, M.; Greeley, J.; Stamenkovic, V.; Markovic, N. M. Nat. Mater. 2012, 11, 550-557.
40
41 (152) Yang, J.; Cooper, J. K.; Toma, F. M.; Walczak, K. A.; Favaro, M.; Beeman, J. W.; Hess, L.
42 H.; Wang, C.; Zhu, C.; Gul, S.; Yano, J.; Kisielowski, C.; Schwartzerg, A.; Sharp, I. D. Nat.
43 Mater. 2017, 16, 335-341.
44
45 (153) Seitz, L. C.; Chen, Z.; Forman, A. J.; Pinaud, B. A.; Benck J. D.; Jaramillo, T. F.
46 ChemSusChem, 2014, 7, 1372-1385.
47
48 (154) Maeda, K.; Teramura, K.; Lu, D.; Saito, N.; Inoue, Y.; Domen, K. Angew. Chem. Int. Ed.
49
2006, 45, 7806-7809.
50
51
(155) Maeda, K.; Domen, K. Top. Curr. Chem. 2011, 303, 95-119.
52
53
54
(156) Yoshida, M.; Takanabe, K.; Maeda, K.; Ishikawa, A.; Kubota, J.; Sakata, Y.; Ikezawa, Y.;
55 Domen, K. J. Phys. Chem. C 2009, 113, 10151-10157.
56
57
58
59
60
ACS Paragon Plus Environment
43
ACS Catalysis Page 44 of 58

1
2
3
(157) Qureshi, M.; Shinagawa, T.; Tsiapis, N.; Takanabe, K. ACS Sustain. Chem. Eng. 2017, 5,
4
5 8079-8088.
6
7 (158) Garcia-Esparza, A. T.; Shinagawa, T.; Ould-Chikh, S.; Qureshi, M.; Peng, X.; Wei, N.;
8 Anjum, D. H.; Clo, A.; Weng, T.-C.; Nordlund, D.; Sokaras, D.; Kubota, J.; Domen, K.;
9 Takanabe, K. Angew. Chem. Int. Ed. 2017, 56, 5780-5784.
10
11 (159) Pan, C.; Takata, T.; Nakabayashi, M.; Matsumoto, T.; Shibata, N.; Ikuhara, Y.; Domen, K.
12 Angew. Chem. Int. Ed. 2015, 54, 2955-2959.
13
14 (160) Yoshida, M.; Maeda, K.; Lu, D.; Kubota, J.; Domen, K. J. Phys. Chem. C 2013, 117,
15
16
14000-14006.
17
18
(161) Hisatomi, T.; Maeda, K.; Takanabe, K.; Kubota, J.; Domen, K. J. Phys. Chem. C 2009,
19 113, 21458-21466.
20
21 (162) Maeda, K.; Xiong, A.; Yoshinaga, T.; Ikeda, T.; Sakamoto, N.; Hisatomi, T.; Takashima,
22 M.; Lu, D.; Kanehara, M.; Setoyama, T.; Teranishi, T.; Domen, K. Angew. Chem. Int. Ed. 2010,
23 49, 4096-4099.
24
25 (163) Townsend, T. K.; Browning, N. D.; Osterloh, F. E. Energy Environ. Sci. 2012, 5, 9543-
26 9550.
27
28 (164) Takanabe, K. In Nanotechnology in Catalysis: Applications in the Chemical Industry,
29
30
Energy Development, and Environment Protection, Sels, B.; Voorde, M.V. Eds. Wiley-VCH:
31 2017; pp. 891-906.
32
33 (165) Rajeshwar, K. J. Phys. Chem. Lett. 2011, 2, 1301-1309.
34
35 (166) Atkins, P.; Paula, J. D. Atkins’ Physical Chemistry, 10th ed., W. H. Freeman and
36 Company, New York, 2014.
37
38 (167) Shinagawa, T.; Takanabe, K. J. Power Sources, 2015, 287, 465-471.
39
40 (168) Haynes, W. M.; Lide, D. R. Handbook of Chemistry and Physics, 92nd ed., CRC Press,
41 Boca Raton, FL, 2011.
42
43 (169) Jin, J.; Walczak, K.; Singh, M. R.; Karp, C.; Lewis, N. S.; Xiang, C. Energy Environ. Sci.
44
45
2014, 7, 3371-3380.
46
47
(170) Chabi, S.; Papadantonakis, K. M.; Lewis, N. S.; Freund, M. S. Energy, Environ. Sci. 2017,
48 10, 1320-1338.
49
50 (171) Kutz, R. B.; Chen, Q.; Yang, H.; Sajjad, S. D.; Liu, Z.; Masel, I. R. Energy Technol. 2017,
51 5, 929-936.
52
53 (172) Klingan, K.; Ringlebm, F.; Zaharieva, I.; Heidkamp, J.; Chernev, P.; Gonzalez-Flores, D.;
54 Risch, M.; Fischer, A.; Dau, H. ChemSusChem 2014, 7, 1301-1310.
55
56 (173) Shinagawa, T.; Takanabe, K. ChemSusChem, 2017, 10, 1318-1336.
57
58
59
60
ACS Paragon Plus Environment
44
Page 45 of 58 ACS Catalysis

1
2
3
(174) Shinagawa, T.; Ng, M. T.-K., Takanabe, K. ChemSusChem DOI: 10.1002/cssc.201701266.
4
5
6
(175) Dionigi, F.; Vesborg, P. C. K.; Pedersen, T.; Hansen, O.; Dahl, S.; Xiong, A.; Maeda, K.;
7 Domen, K.; Chorkendorff, I. Energy Environ. Sci. 2011, 4, 2937-2942.
8
9 (176) Chen, Z.; Jaramillo, T. F.; Deutsch, T. G.; Kleiman-Shwarsctein, A.; Forman, A. J.;
10 Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; McFarland, E. W.; Domen, K.;
11 Miller, E. L.; Turner, J. A.; Dinh, H. N. J. Mater. Res. 2010, 25, 3-16.
12
13 (177) van de Krol, R.; Grätzel, M. Photoelectrochemical hydrogen production, Springer, 2012.
14
15 (178) Esposito, D. V.; Lee, Y.; Yoon, H.; Haney, P. M.; Labrador, N. y.; Moffat, T. P.; Talin, A.
16 A.; Szalai, V. A. Sustainable Energy Fuels, 2017, 1, 154-173.
17
18 (179) Nurlaela, E.; Ziani, A.; Takanabe, K. Mater. Renew. Sustainable Energy, 2016, 5, 18.
19
20 (180) Nurlaela, E.; Ould-Chikh, S.; Harb, M.; Del Gobo, S.; Aouine, M.; Puzenat, E.; Sautet, P.;
21
22 Domen, K.; Basset, J.-M.; Takanabe, K. Chem. Mater. 2014, 26, 4812-4825.
23
24 (181) Ziani, A.; Nurlaela, E.; Dhawale, D. S.; Alves Silva, D.; Alarousu, E.; Mohammed, O. F.;
25 Takanabe, K. Phys. Chem. Chem. Phys. 2015, 17, 2670-2677.
26
27 (182) Nurlaela, E.; Wang, H.; Shinagawa, T.; Flanagan, S.; Ould-Chikh, S.; Qureshi, M.; Mics,
28 Z.; Sautet, P.; Le Bahers, T.; Cánovas, E.; Bonn, M.; Takanabe, K. ACS Catal. 2016, 6, 4117-
29 4126.
30
31 (183) de Respinis, Moreno, Fravventura, M.; Abdi, F. F.; Schreuders, H.; Savenije, T. J.; Smith,
32
W. A.; Dam, B.; van de Krol, R. Chem. Mater. 2015, 27, 7091-7099.
33
34
35
(184) Vequizo, J. J. M. Hojamerdiev, M.; Teshima, K.; Yamakata, A. J. Photochem. Photobiol.
36 A 2017, DOI: 10.1016/j.jphotochem.2017.09.005
37
38 (185) Xing, Z.; Zong, X.; Pan, J.; Wang, L. Chem. Eng. Sci. 2013, 104, 125-146.
39
40 (186) Setoyama, T.; Takewaki, T.; Domen, K.; Tatsumi, T. Faraday Discuss. 2017, 198, 509-
41 527.
42
43 (187) Singh, M. R.; Xiang, C.; Lewis, N. S. Sustain. Energy Fuels 2017, 1, 458-466.
44
45 (188) Pinaud, B. A.; Benck, J. D.; Seitz, L. C.; Forman, A. J.; Chen, Z.; Deutsch, T. G.; James,
46 B. D.; Baum, K. N.; Baum, G. N.; Ardo, S.; Wang, H.; Miller, E.; Jaramillo, T. F. Energy
47 Environ. Sci. 2013, 6, 1983-2002.
48
49
(189) Hisatomi, T.; Miyazaki, K.; Takanabe, K.; Maeda, K.; Kubota, J.; Sakata, Y.; Domen, K.
50
51 Chem. Phys. Lett. 2010, 486, 144-146.
52
53 (190) Hisatomi, T.; Maeda, K.; Takanabe, K.; Kubota, J.; Domen, K. J. Phys. Chem. C 2009,
54 113, 21458-21466.
55
56
57
58
59
60
ACS Paragon Plus Environment
45
ACS Catalysis Page 46 of 58

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
46
Page 47 of 58 ACS Catalysis

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Figure 1. Schematic image of the photocatalytic water splitting process. The gear with the
35 number indicates the order of the photocatalytic process to be successful for overall water
36 splitting. For detailed description, please refer to the text.
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
47
ACS Catalysis Page 48 of 58

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 2. Parameters associated with photocatalysis. Overall water splitting is only successful
27 for high efficiencies of all six gears depicted in the scheme. The different timescales of the
28
29
reactions are also displayed.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
48
Page 49 of 58 ACS Catalysis

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 3. Bandgap structure of oxide and oxynitride semiconductors for photoelectrochemical
22
23
24 applications. Contribution of metal cation and oxygen anion states to the conduction and valence
25
26 bands. The bandgap energy (red for n-type, black for p-type) is shown with respect to the
27
28 reversible hydrogen electrode and the water redox energy levels (assuming Nernstian behavior
29
30
31 four the band-edge energies with respect to electrolyte pH). Reprinted, with permission, from ref
32
33 45.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
49
ACS Catalysis Page 50 of 58

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 Figure 4. (A) Hole and (B) electron lifetimes in heavily doped n-type and p-type silicon,
55
56
57 respectively. Reprinted, with permission, from ref 75.
58
59
60
ACS Paragon Plus Environment
50
Page 51 of 58 ACS Catalysis

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 5. Rough estimation of the ratios of the numbers between the active surface sites
39
40 (assuming ~4 nm−2 hydroxylated surface as maximum)82 to the bulk carrier. The cubic particle of
41
42 100 nm diameter is used as an example.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
51
ACS Catalysis Page 52 of 58

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 6. (A) Barrier height versus electronegativity of metals deposited on Si, GaSe, and SiO2.
26
27
28 (B) Index of interface behavior S as a function of the electronegativity difference of the
29
30 semiconductors. Reprinted, with permission from ref. 89.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
52
Page 53 of 58 ACS Catalysis

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 7. (A) Geometric model schemes using n-type semiconductor with HER catalyst
26
27 decoration with the boundary conditions and the assumptions used for the simulations. (B)
28
29
30 Potential gradients under the HER catalyst (red dotted line in A) at different donor concentration,
31
32 carrier mobility, and carrier lifetime. The x-direction represents the depth from surface (left) into
33
34 the bulk (right) of the semiconductor. An ohmic junction was assumed for the HER catalyst in
35
36
37 contact with the semiconductor, whereas a Schottky contact was assumed to calculate the
38
39 electrolyte interface. The potential difference between HER site and OER site is assumed to be
40
41
42
1.53 eV. Reprinted, with permission, from ref 107.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
53
ACS Catalysis Page 54 of 58

1
2
3
Table 1
4
5
6
List of events, parameters, variables, relevant theories and useful characterization techniques for photocatalysis investigation
7 Events Parameters/ variables Theory Characterization techniques
8
9 1. Photon absorption Absorptance / Reflectance / Scattering Franck-Condon principle X-ray diffraction
10 Absorption coefficient Lambert-Beer’s law UV-VIS-NIR spectroscopy
11 Absorption depth Electromagnetic wave propagation Spectroradiometer
12 Density of state Maxwell Curl equations
2. Exciton separation Effective mass Electrostatic force Transient absorption spectroscopy
13 Dielectric constant / dielectric loss Mott-Wannier type Photoemission spectroscopy
14 Refractive index Frenkel type optical absorption spectroscopy
15 Exciton binding energy photoconductivity screening potential spectroscopy
16 magneto-optical spectroscopy
17 3. Carrier diffusion Carrier mobility Recombination models (srh, Auger) van der Pauw technique with Hall measurement
Diffusion coefficient Poisson equation Time resolved spectroscopy
18
Carrier lifetime Drift and diffusion equations THz and microwave spectroscopies
19 Carrier diffusion length Continuity equations
20 Carrier concentrations Boltzmann transport equation
21 Charge recombination kinetics Semiconductor devices equations
22 4. Carrier transport Electric field Einstein relation Conductivity measurement
23 Drift current Mott-Schottky analysis Photoemission spectroscopy (in air)
Depletion layer width Schottky/ohmic contact Ultraviolet photoemission spectroscopy
24 Flatband potential / workfunction / redox potential Electrochemistry (aqueous non-aqueous)
25 (potential determining ion) Intensity modulated photocurrent/photovoltage
26 Barrier height spectroscopy
27 Fermi level pinning Ambient pressure X-ray photoelectron spectroscopy
28 Density of surface states
Kinetics of charge transfer and recombination
29 5. Electrochemistry Exchange current density (charge transfer resistance) Butler-Volmer analysis Voltammetry, Tafel analysis
30 Charge/electron transfer coefficient Tafel equation Impedance spectroscopy
31 Conductivity
32 Tafel slope
33 Activation energy
6. Mass transfer Diffusion coefficient (ion size viscosity activity Nernst-Planck-Poisson equation Koutechy-Levich analysis
34
coefficient) Fick's law Viscometer
35 Solution resistance Einstein-Smoluchowski equation pH meter
36 Cottrell / Koutechy-Levich equation Conductivity/ impedance
37 Other parameters / Temperature Scanning electron microscope
38 variables and Activity/fugacity (of reactant and products) Transmission electron microscope
39 characterization Photon flux and photon distribution X-ray diffraction
techniques Durability X-ray photoelectron spectroscopy
40
41
42
43
44
45
46 ACS Paragon Plus Environment
47
48 54
Page 55 of 58 ACS Catalysis

1
2
3
Table 2
4
5
6
Major constants and variables involved in photocatalysis
7
8 Symbol Unit Description
9
10 Constants
11
12 e C elementary charge
13
14 kB J K−1 Boltzmann constant
15
16
17 h Js Planck constant
18
19 ε0 F m−1 vacuum permittivity
20
21
22 me kg electron mass
23
24 R J mol−1 K−1 gas constant
25
26
F C mol−1 Faraday constant
27
28
29 Variables
30
31
Semiconductor equations
32
33
34 T K temperature
35
36 εr(s) relative permittivity (dielectric constant) of semiconductor
37
38
39 n, p m−3 electron and hole concentration
40
41 ni m−3 intrinsic carrier concentration
42
43
44 n 0, p 0 m−3 quasi-equilibrium carrier density
45
46 NC, NV m−3 effective density of states in the conduction and valence band
47
48
µ n, µ p m2 V−1 s−1 electron and hole mobility
49
50
51 τn, τp s electron and hole lifetime
52
53 τc s collision time
54
55
56 Dn, Dp m2 s−1 electron and hole diffusion coefficient
57
58
59
60
ACS Paragon Plus Environment
55
ACS Catalysis Page 56 of 58

1
2
3
4 L m diffusion length
5
6 P0 m2 s−1 photons absorbed from AM 1.5G
7
8
9 α(λ) m−1 absorption coefficient
10
11 λ m wavelength of photon
12
13
x m depth into the bulk of a semiconductor
14
15
16 ρ m surface of the semiconductor
17
18 r0, rs m catalyst and semiconductor particle size (diameter)
19
20
21 χ eV semiconductor electron affinity
22
23 Eg eV band gap
24
25
26 EC eV conduction band edge
27
28 EV eV valence band edge
29
30
m *n, m *p effective electron and hole mass
31
32
33 A *n, A *p A m−2 K−2 effective Richardson constant for electrons and holes
34
35 Electrochemical parameters
36
37
38 n number of electrons in reaction
39
40 ai thermodynamic activity (of species i)
41
42
43 γ± activity coefficient
44
45 Di m2 s−1 diffusion coefficient (of species i)
46
47
48 δ m diffusion layer thickness
49
50 u m2 s−1 V−1 ion mobility
51
52
a m Stokes radius
53
54
55 µ Pa s viscosity of solution
56
57
58
59
60
ACS Paragon Plus Environment
56
Page 57 of 58 ACS Catalysis

1
2
3
4 ν m2 s−1 kinematic viscosity of solution
5
6 εr(l) relative permittivity (dielectric constant) of solution
7
8
9 η V overpotential
10
11 α transfer coefficient
12
13
j0 A cm−2 exchange current density
14
15
16 θ surface coverage
17
18 (depending on
k rate constant
19 elementary steps)
20
(depending on
21 A preexponential factor
elementary steps)
22
23
Ea kJ mol−1 activation energy
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
57
ACS Catalysis Page 58 of 58

1
2
3
TOC
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
58

You might also like