Surface Localization of Defects in Black TiO2 ... - ACS Publications

Dec 8, 2016 - Despite the excellent optical and electrical features of black TiO2 for ... Interfacial Engineering Using Ag-doped TiO 2 Hole Blocking L...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Otago Library

Perspective

Surface Localization of Defects in Black TiO2: Enhancing Photoactivity or Reactivity Kan Zhang, and Jong Hyeok Park J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.6b02289 • Publication Date (Web): 08 Dec 2016 Downloaded from http://pubs.acs.org on December 9, 2016

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.

The Journal of Physical Chemistry Letters 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 33

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

The Journal of Physical Chemistry Letters

Surface Localization of Defects in Black TiO2: Enhancing Photoactivity or Reactivity Kan Zhang*, Jong Hyeok Park* †

Department of Chemical and Biomolecular Engineering, Yonsei University, 50 Yonsei-ro, Seodaemun-

gu, Seoul 120-749, Republic of Korea

Corresponding Author

K.Z ([email protected]), J. H. Park ([email protected])

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry Letters

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

Page 2 of 33

ABSTRACT In the past several years, surface-disordered TiO2, which is referred to as black TiO2 and can absorb both visible and near-infrared solar light, has triggered an explosion of interest for many important applications. Despite the excellent optical and electrical features of black TiO2 for various photoelectrochemical (PEC) and photochemical reactions, the current understanding of the photocatalytic mechanism is unsatisfactory and incomplete. Based on previous studies, we present new insight into the surface localization of defects and perspectives on the liquid/solid interface. The future prospects for understanding black TiO2 from this perspective suggest that defect engineering at the liquid/solid interface is a potential method of guiding nanomaterial design.

TOC GRAPHIC

KEYWORDS: Black TiO2, disorder, photoactivity, reactivity, interface engineering

ACS Paragon Plus Environment

2

Page 3 of 33

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

The Journal of Physical Chemistry Letters

Titanium dioxide (TiO2), which is also called titanium pigment, is widely used as an additive in sunscreen cream because it can efficiently absorb the ultraviolet (UV) light in sunlight. In 1972, a pioneering work by Fujishima and Honda demonstrated that crystalline TiO2 can convert absorbed UV light into a charge flow to split water [1], opening up a wide range of potential eco-friendly applications for solar in chemical fuel conversion and environmental purification [2-5]. In particular, the most stable TiO2 has an excellent band structure that can produce electron/hole pairs with a high redox ability to drive numerous photocatalytic reactions. It is not, however, suitable for practical applications because of several disadvantages. The main challenges are the limited charge flow under sunlight illumination because of its large band gap (3.2 eV) and its rapid charge recombination. Because only ~4% of the solar spectrum lies in the UV region, the solar energy conversion efficiency of TiO2 remains impractical. Therefore, tremendous efforts have been expended in the past decade to harvest visible photons because visible light accounts for ~46% of sunlight illumination [6-8]. It is commonly believed that the electronic structure of a semiconductor is intimately associated with its band gap—specifically, its light-harvesting region—and its subsequent photocatalytic performance [9,10]. In many cases, however, although a suitable value of the band gap is beneficial for light absorption, the reactivity is closely related to several other bulk and surface properties. Indeed, the controlling bulk (charge transport) and surface (chemical reaction) recombination processes are critical for photocatalytic performance (Figure 1).[11] Relatively far-sighted investigations have indicated that these properties of TiO2, including its reactivity and photoreactivity, are strongly related to its defect disorder.[12,13] This is especially true in the defect engineering of TiO2, in which crystal and surface defects are produced by oxygen vacancy, Ti3+ doping or atomic arrangement disorder and can profoundly influence the photoreactivity in terms of its electronic structure,

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry Letters

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

Page 4 of 33

charge transport, and surface properties.[14, 15]

Figure 1. Schematics of bulk and surface charge recombination during photocatalysis.

In 2011, a black TiO2 produced by hydrogenation was reported [16]. Unsurprisingly, the dramatic color change from white to black involves near-infrared light absorption (Figure 2a and b). In contrast to cases of doping TiO2 that exhibited various colors because of the introduction of new band levels, the band gap narrowing of black TiO2 was ascribed to a thin disordered layer surrounding ordered TiO2, as shown in Figure 2c and d [17, 18]. However, as noted by Dr. Diebold, the finding that introducing a disordered surface layer in a TiO2 crystal yields better photocatalytic ability is contrary to conventional wisdom [19]. To date, despite various efforts to study defective TiO2 through reduction under various conditions, controversy persists regarding the real role of disordered TiO2 in photocatalysis. On the one hand, the presence of the disordered part narrows the bandgap of TiO2, allowing near-infrared light absorption; however, the photoactivity of black TiO2 does not match its attractive absorption spectra [20, 21]. On the other hand, another issue—whether the disordered part modifies the surface properties of black TiO2, tuning its reactivity reactive with water or other substances—remains unaddressed.

ACS Paragon Plus Environment

4

Page 5 of 33

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

The Journal of Physical Chemistry Letters

However, these properties are not strongly supported by this material’s photocatalytic activity. In the past five years, studies on defective TiO2 have continued, and progress has been made on elucidating its chemical composition, electronic structure, and photoactivity mechanisms and in developing new applications. This perspective is intended to bring together the photocatalytic and photoelectrochemical (PEC) aspects to understand the superior reactivity and photoactivity of black TiO2.

Figure 2. a) Pictures of white and black TiO2 nanocrystals. b) Optical absorption spectra of white and black TiO2. Reproduced with permission.[16] Copyright 2011, American Association for the Advancement of Science. High-resolution transmission electron microscopy images: (c) white TiO2 and (d) black TiO2 nanocrystal. Reproduced with permission.[17] Copyright 2013, Nature Publishing Group.

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry Letters

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

Page 6 of 33

Do the atomic defects in black TiO2 significantly improve the photocatalytic performance? Atomic defects, which are mainly caused by oxygen vacancies or Ti3+ species, were previously believed to enhance the photocatalytic performance through bandgap narrowing or charge separation. These atomic defects are generally induced via two approaches: inherent oxygen deficiency resulting from the pre-formed synthesis conditions and post-treated TiO2 under reducing conditions. The black TiO2 obtained by high-pressure hydrogen treatment is naturally associated with such atomic defects [22]. However, the formation of oxygen vacancies or Ti3+ species upon TiO2 reduction is supported by several pieces of experimental evidence. (1) The color is tunable between yellow and blue, which is ascribed to d-d transitions.[23] (2) The conduction band (CB) minimum is decreased by ∼1 eV, and the downward CB position usually acts as an electron acceptor because of localized electronic states (Figure 3a) [24,25]. These two characteristics associated with atomic defects were inconsistent with black TiO2, which shows a maximum valance band (VB) upward shift of 2.18 eV [16]. Generally, the upward VB position for the formation of the mid-gap state results from a band of delocalized states (Figure 3b) [26, 27] through which multiphoton excitation might be achieved. However, black TiO2 has only a certain degree of photoactivity under visible light that is similar to the photocatalytic performance gain contributed by oxygen vacancies or Ti3+ species in TiO2 [20, 28]. However, no direct evidence regarding the long-wavelength effects on black TiO2 was reported until recently. Irrespective of light absorption, black TiO2 exhibits much higher photoactivity than oxygen vacancy-containing or Ti3+-defective TiO2. Subsequent investigations involving black TiO2 revealed that the main contributor to its photoactivity is still likely to be oxygen vacancies, which improve charge separation and whole carrier density [29-31]. However, this conclusion does not fully explain why the hydrogenated TiO2 is black or the reason underlying its impressive

ACS Paragon Plus Environment

6

Page 7 of 33

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

The Journal of Physical Chemistry Letters

photoactivity beyond the effects of oxygen vacancies or Ti3+ species. Naldoni and co-workers proposed that black TiO2 results from a synergistic effect of the oxygen vacancies on core and surface disorder [32]. Conversely, Tian et al. attributed black TiO2’s characteristics to the surface formation of a Ti2O3 shell accompanied by lattice distortion in the core [33]. Although both explanations for the formation of black TiO2 were consistent with the electronic structure, whether its high photoactivity also originates from these synergistic effects remains unclear. The photocatalytic performance and electronic structure of defective TiO2 relies strongly on the concentration and distribution of oxygen vacancies or Ti3+ [34]. For example, Tan et al. performed a detailed bandgap alteration of atomically defective TiO2 with a gradually increasing concentration of oxygen vacancies or Ti3+ based on first-principles density functional theory (DFT) calculations, demonstrating a downward band shift and the development of multiple bands with widths 0.5-1.5 eV below the CB minimum [35]. Kong et al. observed that decreasing the relative concentration ratio of bulk defects to surface defects in TiO2 nanocrystals significantly improves the charge separation efficiency and, thus, enhances the photoactivity.[13] As a result, the superior photoactivity of black TiO2 can be tentatively attributed to the surface localization of defects, leading users to investigate trap states and reactivity. Leshuk et al. reported the hydrogenation of TiO2 at temperatures ranging from 250 to 450 °C under atmospheric pressure. All the samples showed worse photoactivity than the unhydrogenated one, allowing conclusions to be drawn regarding the excessive formation of bulk defects.[36] Yu et al. used similar synthesis conditions for the hydrogenation of TiO2 and found that bulk defects diffuse to the surface at temperatures above 600 °C, thereby greatly improving the photoactivity [37].

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry Letters

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

Page 8 of 33

Figure 3. Excitation state of oxygen vacancy-containing or Ti3+-defective TiO2 (a) and black TiO2 (b). Furthermore, in addition to the photochemical reactions of black TiO2, this material’s PEC performance demonstrates significantly enhanced conductivity and carrier density [38-40]. The characterized properties are exactly the same as those of atomically defective TiO2 (oxygen vacancies or Ti3+ doping) in which the conductivity and carrier density are also increased [4146]. However, another electrochemical investigation revealed that the surface disordered layer can serve as a passivation layer that enhances the surface water oxidation kinetics [47, 48] and stability [49]. In our recent work, we systemically investigated the surface properties of black TiO2 in light of its PEC properties and demonstrated the remarkable narrowing of the depletion layer and increasing electron diffusion length, which facilitated unlimited electron collection in a 20-µm length of TiO2 (Figure 4) [50]. Because the photocatalytic reaction is included among the surface thermodynamics reactions, the findings regarding the solid/liquid interface demonstrate the unusual photoactivity and reactivity of black TiO2, which affect the overall photocatalytic performance in a different manner than simple atomic defects in the bulk material. On the other

ACS Paragon Plus Environment

8

Page 9 of 33

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

The Journal of Physical Chemistry Letters

hand, if the disorder layer could generate electron/hole pairs under light illumination, the open circuit voltage of PEC device should be changed. Unfortunately, the classic phenomenon regarding bandgap narrow in form of heterojunction could not be reported as far as we known [50].

Figure 4. (a) Schematic of the configuration of black TiO2 wire arrays. (b) Plots of the photocurrent densities (1.23 V vs reversible hydrogen electrode [RHE]) and enhancement factors as a function of the length of the wire arrays. Energy diagram for the charge-transfer equilibrium with the electrolyte: (c) TiO2 and (d) black TiO2. Reproduced with permission.[50] Copyright 2014, Wiley.

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry Letters

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

Page 10 of 33

Does surface disorder in black TiO2 significantly improve its photocatalytic performance? The photocatalytic process involves three crucial steps: light absorption, charge separation in the bulk region, and charge recombination on the surface. The first two processes have received the most attention, with much less attention being focused on the third. Presumably, this step is taken for granted after loading the co-catalyst. Our recent work developed a Janus crystalline/disordered TiO2 nanoparticle with an open-ended interface that exhibited impressive photocatalytic H2 activity with no co-catalyst (Figure 5a). The time-correlated single-photon counting (TCSPC) results indicated that the Janus crystalline/disordered TiO2 nanoparticle efficiently suppresses indirect electron/hole pair recombination at surface traps or defects [51]. The Schmuki group also developed black TiO2 via hydrogenation or high-energy proton implantation and claimed that the disordered region of this material acts as a catalytic center for co-catalyst-free H2 evolution (Figure 5b) [52,53]. In principle, from a catalytic perspective, the presence of structural defects helps to decrease the coordination number of the neighboring sites on the material surface, thereby tailoring the electronic structure and catalytic activities [54]. Therefore, both of the above findings seem to allow co-catalyst-free H2 evolution from black TiO2 because both the surface trap state and the catalytic center are crucial for the evolution of H2 from water on the photocatalyst surface. In the previous two decades, surface defects (i.e., both Ti3+ and oxygen vacancies) were demonstrated to spontaneously cause water dissociation by forming a metastable surface active complex, resulting in high reactivity [55-57]. In contrast, the surface disordered layer efficiently covered the intrinsic trapping states of TiO2, creating an opportunity for water redox outside of other intermediaries [58, 59]. However, the co-catalytic activity of the disorder part has not yet been confirmed.

ACS Paragon Plus Environment

10

Page 11 of 33

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

The Journal of Physical Chemistry Letters

Figure 5. (a) Comparison of the co-catalyst-free H2 generation and cycling performance of P25 and blue P25 after 1 day of continuous reaction using methanol as a sacrificial agent. (b) Photocatalytic H2 production under open-circuit conditions in methanol/water (50/50 vol %) with TiO2 nanotubes and nanorods treated in different atmospheres under AM 1.5 (100 mW/cm2) illumination. Reproduced with permission.[51] Copyright 2016, The Royal Society of Chemistry. Reproduced with permission.[52] Copyright 2015, American Chemical Society. Similar performance gains in terms of both PEC water splitting and photocatalytic H2 generation have been achieved using a distinct thin amorphous TiO2 layer [60-63]. The surface amorphous layer can passivate surface states at the electrode/electrolyte interface or affect the space-charge layer width to decrease electron-hole recombination pathways. Generally, amorphous materials exhibit a set of universal properties that are distinct from those of their crystalline counterparts. A crystal can evolve away from perfect crystalline order and eventually develop into an amorphous state when disorder is introduced [64, 65]. However, how disorder causes such deviation in the properties of a solid from those of a crystal remains unclear; indeed, no standardized model exists for theoretical simulations of the disordered structure [19]. In contrast to the thinner amorphous layer that usually shields light penetration from reaching the

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry Letters

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

Page 12 of 33

active core materials, this disordered layer unexpectedly showed gradually improved light absorption as its thickness increased [66-71]. These studies indicate that the disordered surface in black TiO2 may have unusual properties that affect the bulk material. Charge transport between the ordered and disordered parts in black TiO2 Although increasing interest has developed regarding the photochemical or PEC properties of black TiO2 over the past five years, these properties have been attributed to the prominent charge separation efficiency in the bulk region. By performing calculations based on DFT and the Perdew-Burke-Ernzerhof (PBE) functional, Liu et al. demonstrated that the charge separation mechanism in black TiO2 is based on the localization of holes accompanied by the delocalization of electrons [18]. The wave function of itinerant conduction electrons spreads throughout both the crystalline and disordered regions, but the holes are localized in the midgap state of the disordered part. The photo-induced charge generation and separation in black TiO2 are treated as integral behaviors, regardless of their respective occurrence in the crystalline or disordered part. Our recent work revealed spatially separated crystalline and disordered regions in Janus TiO2 and determined the deteriorated charge separation in fully disordered TiO2. However, rapid charge separation could be confirmed when the disordered TiO2 was placed in heterogeneous contact with crystalline TiO2 [50]. The results clearly showed the strong synergistic effect on charge transfer between the crystalline and disordered parts. The Xu and Jia group focused on interfacial charge transfer at the crystalline and disorder junction and confirmed that unparalleled metallic conduction is possible at their interface [72]. To date, however, the alignment of charge transfer routes with type II or the movement of these routes from the crystalline region to the disordered region has not been observed. In addition, whether the electronic states of the Ti3+ species are localized or delocalized remains unknown, whereas the conductivity and charge

ACS Paragon Plus Environment

12

Page 13 of 33

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

The Journal of Physical Chemistry Letters

carrier density of black TiO2 have been shown to be orders of magnitude higher than those of crystalline TiO2 [35, 73]. Consequently, once a charge is transferred to the disordered part, its superior conductivity will accelerate charge transport. Contradicting concerns Because of its unprecedented texture and unusual electrochemical and optical properties, black TiO2 has been the subject of substantial research activities investigating its application in photocatalytic or PEC water splitting. However, its formation and photocatalytic mechanism remain somewhat controversial. Chen et al. attributed the formation of black TiO2 to H doped into the TiO2 lattice with H-Ti bonds. Danon et al. suggested that the black color of hydrogenated TiO2 originates from the doping of Cr ion derived from the stainless steel reactor [74]. Indeed, hydrogenated TiO2 was found to be blue when created in a quartz reactor instead of a stainless steel reactor. Other explanations proposed for the black color include the formation of titanium sub-oxides, such as Ti2O3 [33], Ti4O7 [75], or other Magnéli phase crystals [76] via strong reduction. Another issue is the photocatalytic mechanism of black TiO2 with a core/shell structure. Photocatalysis mainly occurs at the surface, which requires not only efficient electron/hole separation in the bulk but also charge transport to surface redox sites. Zhu et al. demonstrated that the photocatalytic H2 performance of platinized black TiO2 is much worse than that of hydrogenated platinized TiO2 [77]. These results suggest that the surface disorder involves confined itinerant electrons migrating to active sites of the TiO2 surface to some extent, although the disordered surface is subsequently loaded with Pt. Nandasiri et al. demonstrated that the H implantation into (110) rutile TiO2 lattice is detrimental to its stability because of the outward diffusion of H atoms, in addition to desorption and surface reduction [78], which contradicts

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry Letters

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

Page 14 of 33

other reports. Does other possibility of surface textures affect its photoactivity and reactivity? Actually, the surface reactivity of TiO2 was triggered a lot interest for a long time. The earlier reviews titled “photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results” and “surface science studies of the photoactivation of TiO2-new photochemical processes” have emphasized the importance of TiO2 surface chemistry [79, 80]. In particular, the reactivity of TiO2 caused by the surface defects, such as bonding energies with molecule, chemisorbed energy and hydrophilicity are an element of coherence in understanding. Most of the surface reactivity on TiO2 have been concerned with various theory and practice, their photocatalytic and photochemical activities were proved to associate with sub-band-gap bulk and surface electronic states of TiO2 [81-87]. In consequence, the surface properties are closely related to TiO2 polymorphs. So, what is the surface reactivity of disorder layer when the disorder layer formed on the surface of different TiO2 polymorphs? Liu et. al investigated the co-catalytic activities of hydrogenated black TiO2 with different polymorphs (anatase, rutile, anatase/rutile). Interestingly, as shown in Figure 5a, the disorder layer formed on rutile TiO2 surface could not play cocatalytic center for H2 generation, whereas the disorder layer on anatase/rutile TiO2 junction can generate a strong synergistic effect for co-catalytic H2 generation [88]. Therefore, the defect configurations in disorder layer for different TiO2 polymorphs seems to be crucial. As Chen et al pointed out, the photocatalytic activity of black TiO2 relied on its preparation method, just like the co-catalytic H2 activity induced by oxygen vacancies, Ti3+ states, Ti interstitials, and surface reconstruction could not observed for black TiO2 reduced by other common methods [21]. Xia and Chen studied the structural properties of hydrogenated black TiO2 from the percentage of various crystalline facets and the ratio of the crystalline/amorphous phases aspects,

ACS Paragon Plus Environment

14

Page 15 of 33

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

The Journal of Physical Chemistry Letters

demonstrating that the surface excess pressure and stress induced by disorder layer have a different exposed crystalline facet tendency from the thermodynamically stable order [89]. In line with other electrocatalysts [90, 91], the strain-texture can tune surface electronic potential, which is able to be responsible for catalytic activity. However, to draw an exact conclusion on the surface defect inducing active sites, some in-situ analysis tools targeting catalytic hotspots are expected, for example, scanning electrochemical microscopy, super-resolution fluorescence imaging et al. Liu’s report lefts behind another question that is how the black TiO2 comprising mixed anatase and rutile phases can produce high reactivity. According to theoretical simulation by Deskins’s group, the unpaired electrons are more stable within the interfacial region than in the pure phases alone (Figure 5b). As a result, the interfacial region is potentially more reactive, since localized electrons could be transferred to adsorbates on the surface near the interfaces [92]. For the black TiO2 with mixed anatase and rutile phases, the co-catalytic center located at their interface, namely a disorder layer at anatase and rutile interface, will produce a higher reactivity for co-catalyst free H2 generation of TiO2 or other applications. Xia et al reported that the black TiO2 nanocrystals with mixed anatase and rutile phase can induce excellent microwave absorption

performance,

due

to

the

collective-movement-of-interfacial-dipole

in

crystalline/disordered and anatase/rutile interfaces [93]. The results also suggested that the disorder layer could induce an interfacial polarization effect by the presence of defects (Figure 5c). The located defects in black TiO2 surface are therefore proposed to form interface dislocations or new grain boundaries, electronic defects or phase boundaries. As above mentioned, the disordered TiO2 might be different from amorphous one, a simple experiment can be employed to determine the fact. It is well-known that thermally converting

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry Letters

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

Page 16 of 33

amorphous TiO2 to anatase crystalline is more favourable than converting that to rutile TiO2 at low temperature. When a lower sintering process (200 °C) was performed to crystallize fully disordered rutile TiO2 in mixed anatase/rutile crystal, the obtained production still was rutile TiO2, rather than anatase TiO2 (Figure 5d), which is a direct evidence for the difference between amorphous and disorder crystal. So, the atomic structure of the disordered TiO2 needs to be taken into account which should have a special atomic arrangement differing from amorphous TiO2. Then, an elaborate structure characterization in atomic scale should be carried out to understand the ill-defined geometry. Scanning tunneling microscope (STM) or scanning transmission electron microscopy (STEM) could be considered as currently suitable surface analysis tools to define these unexplored atomic arrangements. Furthermore, the first-principles calculations on photoactivity and reactivity can be set up based on geometry model obtained by STM or STEM. An accurate theoretical model for the black TiO2 is being expected to give the right direction for surface engineered photocatalyst as well as recently pointing out electrocatalyst in the clean fuel generation.

ACS Paragon Plus Environment

16

Page 17 of 33

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

The Journal of Physical Chemistry Letters

Figure 6. (a) Rates of photocatalytic H2 evolution from various TiO2 particle suspensions under AM 1.5 (100 mWcm2) illumination. a) Comparison of different anatase (an, an1, an2), rutile (ru), and mixed-phase (m1, m2) powders before and after hydrogenation (Hy-) at 500 °C,20 bar (powder Ar/H2 exposed to Ar/H2 stream at 500 °C for 1h; powder Ar exposed Ar steam at 500 °C for 1h). Reproduced with permission.[80] Copyright 2014, Wiley. (b) Electron spin density plot (isovalue = 0.017e/A3) of the interfacial region for a slab with vacuum-exposed surfaces. The results are from a rutile (001)/anatase (100) interface with U = 8.4/6.3 eV for rutile/anatase. Reproduced with permission.[92] Copyright 2015, AIP Publishing LLC. (c) The collective movements of interfacial dipoles at the anatase/rutile and crystalline/disordered interfaces amplify the response to the incoming electromagnetic field and thus induce enhanced microwave absorption performance. Reproduced with permission.[93] Copyright 2013, Wiley. (d) XRD pattern of P25 (mixed anatase/rutile TiO2), blue P25 with disordered rutile TiO2 (prepared by ref. 51) and re-annealed blue P25.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry Letters

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

Page 18 of 33

Perspective regarding the surface localization of defects Black TiO2 has been the subject of extensive scientific research seeking to understand its chemical composition and electronic structure and the origin of its photocatalytic activity. Although some issues that may affect its functions remain unaddressed, the high conductivity and narrow bandgap of black TiO2 favor numerous applications relating to electrochemistry, such as batteries [94-97], supercapacitors [98, 99], fuel cells [100] solar cells [101, 102] and PEC sensors [73]. However, the foundational principles of the above electrochemical devices regarding the photoactivity and reactivity at solid/liquid interface have been ignored. Black TiO2 containing enriched defects at the surface is essentially the same as black TiO2 with other surface modifications [103]. However, the disordered texture caused by the surface localization of defects clearly distinguished this material from its crystalline and amorphous counterparts, especially when it is in intimate contact with liquid electrolyte or reactant. An accurate structure characterization in atomic scale combined with relevant theoretical simulation are therefore expected to promote our in-depth fundamental understanding on surface localization of defects for black TiO2. The surface defect or several nearest subsurface layer could be qualitatively analyzed by X-ray absorption near edge structure (XANES) and X-ray absorption fine structure (XAFS) to probe local structural details around specific metal atoms and are sensitive to the oxidation states of the metal [104, 105]. An atomically layered nanomaterials with disordering can be proposed as an ideal platform for the surface analysis and electron microscopy and also can be realistic model for theoretical simulation. We therefore appeal the material scientist, surface chemist and physico-chemist to pay their attention to this respect. Further research on black TiO2 should involve interface characterization during various electrochemical reactions rather than focusing exclusively on optical and electrical properties. The final objective should

ACS Paragon Plus Environment

18

Page 19 of 33

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

The Journal of Physical Chemistry Letters

be to modulate interfacial reactions by engineering disorder so that this material can be used in numerous energy-related applications.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry Letters

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

Page 20 of 33

ASSOCIATED CONTENT AUTHOR INFORMATION Kan Zhang is currently a research professor in the Department of Chemical and Biomolecular Engineering, Yonsei University, Republic of Korea, and was a postdoctoral research fellow at the same institution. He obtained his Ph.D. degree from the SKKU Advanced Institute of Nanotechnology (SAINT) at Sungkyunkwan University, Republic of Korea, in 2015, and received the “Outstanding Self-Financed Students Abroad (Korea)” award in 2013 from the Chinese government. His research interests involve the synthesis and modification of metal oxides and sulfides for solar energy conversion and electrochemical energy storage. Jong Hyeok Park is an associate professor in the Department of Chemical and Biomolecular Engineering, Yonsei University, Republic of Korea. He received his Ph.D. in chemical engineering from KAIST, Republic of Korea, in August, 2004. He then joined the University of Texas at Austin, TX, USA, as a postdoctoral researcher in 2004 (under Prof. Allen J. Bard). From 2008 to 2014, he was an associate professor at Sungkyunkwan University, Republic of Korea. He is the author or co-author of 200 papers and 50 patents.

Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS This work was supported by the NRF of Korea Grant funded by the Ministry of Science, ICT, and

Future

Planning

(NRF-2013R1A2A1A09014038,

2015M1A2A2074663,

ACS Paragon Plus Environment

20

Page 21 of 33

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

The Journal of Physical Chemistry Letters

2016M3D3A1A01913254 (C1 Gas Refinery Program)). This work was supported in part by the Yonsei University Future-leading Research Initiative of 2015(2015-22-0067).

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry Letters

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

Page 22 of 33

REFERENCES [1] Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature. 1972, 238, 37-38. [2] Ma, Y.; Wang, X. L.; Jia, Y. S.; Chen, X. B.; Han, H. X.; Li, C. Titanium Dioxide-Based Nanomaterials for Photocatalytic Fuel Generations. Chem. Rev., 2014, 114, 9987-10043. [3] Hoffmann, M. R.; Martin, S. T.; Choi, W. Y.; Bahnemann, D. W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69-96. [4] Bard, A. L.; Fox, M. A. Artificial photosynthesis: Solar splitting of water to hydrogen and oxygen. Acc. Chem. Res. 1995, 28, 141-145. [5] Ochiai, T.; Fujishima, A. Photoelectrochemical properties of TiO2 photocatalyst and its applications for environmental purification. J. Photochem. Photobio. C. 2012, 13, 247-262. [6] Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253-278. [7] Maeda, K.; Domen, K. Photocatalytic Water Splitting: Recent Progress and Future Challenges. J. Phys. Chem. Lett. 2010, 1, 2655-2661. [8] Tong, H.; Ouyang, S. X.; Bi, Y. P.; Umezawa, N.; Oshikiri, M.; Ye, J. H. Nanophotocatalytic Materials: Possibilities and Challenges. Adv Mater. 2012, 24, 229-251. [9] Guo, Y. Q.; Xu, K.; Wu, C. Z.; Zhao, J. Y.; Xie, Y. Surface chemical-modification for engineering the intrinsic physical properties of inorganic two-dimensional nanomaterials. Chem. Soc. Rev. 2015, 44, 637-646. [10] Liu, L.; Chen, X. B. Titanium Dioxide Nanomaterials: Self-Structural Modifications. Chem. Rev. 2014, 114, 9890-9918. [11] Nowotny, J.; Sheppard, L. R.; Bak, T.; Nowotny, M. K. Titanium dioxide for solar-

ACS Paragon Plus Environment

22

Page 23 of 33

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

The Journal of Physical Chemistry Letters

hydrogen II. Defect chemistry. Int. J. Hydrogen Energy. 2007, 32, 2630-2643. [12] Li, L. D.; Yan, J. Q.; Wang, T.; Zhao, Z. J.; Zhang, J.; Gong, J. L.; Guan, N. J. Sub-10 nm rutile titanium dioxide nanoparticles for efficient visible-light-driven photocatalytic hydrogen production. Nat. Commun. 2015, 6, 5881. [13] Kong, M.; Li, Y. Z.; Chen, X.; Tian, T. T.; Fang, P. F.; Zheng, F.; Zhao, X. J. Tuning the Relative Concentration Ratio of Bulk Defects to Surface Defects in TiO2 Nanocrystals Leads to High Photocatalytic Efficiency. J. Am. Chem. Soc. 2011, 133, 16414-16417. [14] Nowotny, M. K.; Sheppard, L. R.; Bak, T.; Nowotny, J. Defect Chemistry of Titanium Dioxide. Application of Defect Engineering in Processing of TiO2-Based Photocatalysts. J. Phys. Chem. C. 2008, 112, 5275-5300. [15] Thompson, T. L.; Yates Jr. J. T. Surface Science Studies of the Photoactivation of TiO2New Photochemical Process. Chem. Rev. 2006, 106, 4428-4453. [16] Chen, X. B.; Liu, L.; Yu, P. Y.; Mao, S. S. Increasing Solar Absorption for Photocatalysis with Black Hydrogenated Titanium Dioxide Nanocrystals. Science. 2011, 331, 746-750. [17] Chen, X. B.; Liu, L.; Liu, Z.; Marcus, M. A.; Wang, W. C.; Oyler, N. A.; Grass, M. E.; Mao, B.; Glans, P. A.; Yu, P. Y.; Guo, J.; Mao, S. S. Properties of Disorder-Engineered Black Titanium Dioxide Nanoparticles Through Hydrogenation. Sci. Rep. 2013, 3, 1510. [18] Liu, L.; Yu, P. P.; Chen, X.; Mao, S. S.; Shen, D. Z. Hydrogenation and Disorder in Engineered Black TiO2. Phys. Rev. Lett. 2013, 111, 065505. [19] Diebold, U. Photocatalysts, Closing the gap. Nat. Chem. 2011, 3, 271-272. [20] Hu, Y. H. A highly efficient photocatalyst-hydrogenated black TiO2 for the photocatalytic splitting of water. Angew. Chem. Int. Ed. 2012, 51, 12410-12412. [21] Chen, X. B. Liu, L.; Huang, F. Q. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 1861-1885.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry Letters

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

Page 24 of 33

[22] Zhou, W.; Li, W.; Wang, J. Q.; Qu, Y.; Yang, Y.; Xie, Y.; Zhang, K. Wang, L.; Fu, H. G.; Zhao, D. Y. Ordered Mesoporous Black TiO2 as Highly Efficient Hydrogen Evolution Photocatalyst. J. Am. Chem. Soc. 2014, 136, 9280-9283. [23] Khomenko, V. M.; Langer, K.; Rager, H.; Fett, A. Electronic absorption by Ti3+ ions and electron delocalization in synthetic blue rutile. Phys. Chem. Miner. 1998, 25, 338-346. [24] Henrich, V. E.; Dresselhaus, G.; Zeiger, H. J. Observation of Two-Dimensional Phases Associated with Defect States on the Surface of TiO2. Phys. Rev. Lett. 1976, 36, 1335. [25] Kurtz, R. L.; Stock-Bauer, R.; Madey, T. E. Synchrotron radiation studies of H2O adsorption on TiO2 (110). Surf. Sci. 1989, 218, 178-200. [26] Serpone, N.; Emeline, A. V. Semiconductor Photocatalysis-Past, Present, and Future Outlook. J. Phys. Chem. Lett. 2012, 3, 673-677. [27] Mocatta, D.; Cohen, G.; Schattner, J.; Millo, O.; Rabani, E.; Banin, U. Heavily Doped Semiconductor Nanocrystal Quantum Dots. Science. 2011, 332, 77-81. [28] Liu, X.; Zhu, G.; Wang, X.; Yuan, X.; Lin, T.; Huang, F. Q. You have full text access to this content Progress in Black Titania: A New Material for Advanced Photocatalysis. Adv. Energy. Mater. 2016, 6, 1600452. [29] Jiang, X. D.; Zhang, Y. P.; Jiang, J.; Rong, Y. S.; Wang, Y. C.; Wu, Y. C.; Pan, C. X. Characterization of Oxygen Vacancy Associates within Hydrogenated TiO2: A Positron Annihilation Study. J. Phys. Chem. C. 2012, 116, 22619-22624. [30] Pesci, F. M.; Wang, G. M.; Klug, D. R.; Li, Y.; Cowan, A. J. Efficient Suppression of Electron–Hole Recombination in Oxygen-Deficient Hydrogen-Treated TiO2 Nanowires for Photoelectrochemical Water Splitting. J. Phys. Chem. C. 2013, 117, 25837-25844. [31] Rex, R.; Yang, Y.; Knorr, F.; Zhang, J. Z.; Li, Y.; McHale, J. Spectroelectrochemical

ACS Paragon Plus Environment

24

Page 25 of 33

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

The Journal of Physical Chemistry Letters

Photoluminescence of Trap States in H-Treated Rutile TiO2 Nanowires: Implications for Photooxidation of Water. J. Phys. Chem. C. 2016, 120, 3530-3541. [32] Naldoni, A.; Allieta, M.; Santangelo, S.; Marelli, M.; Fabbri, F.; Cappelli, S.; Bianchi, C. L.; Psaro, R.; Santo, V. D. Effect of Nature and Location of Defects on Bandgap Narrowing in Black TiO2 Nanoparticles. J. Am. Chem. Soc. 2012, 134, 7600-7603. [33] Tian, M. K.; Mahjouri-Samani, M.; Eres, G.; Sachan, R.; Yoon, M.; Chisholm, M. F.; Wang, K.; Puretzky, A. A.; Rouleau, C. M.; Geohegan, D. B.; Duscher, G. Structure and Formation Mechanism of Black TiO2 Nanoparticles. ACS Nano. 2015, 9, 10482-10488. [34] Henderson, M. A. Evidence for bicarbonate formation on vacuum annealed TiO2 (110) resulting from a precursor-mediated interaction between CO2 and H2O. Surf. Sci. 1998, 400, 203219. [35] Tan, H. Q.; Zhao, Z.; Niu, M.; Mao, C. Y.; Cao, D. P.; Cheng, D. J.; Feng, P. Y.; Sun, Z. C. A facile and versatile method for preparation of colored TiO2 with enhanced solar-driven photocatalytic activity. Nanoscale. 2014, 6, 10216-10223. [36] Leshuk, T.; Parviz, R.; Everett, P.; Krishnakumar, H.; Varin, R. A. Photocatalytic Activity of Hydrogenated TiO2. ACS Appl. Mater. Interfaces. 2013, 5, 1892-1895. [37] Yu, X. M.; Kim, B. S.; Kim, Y. K. Highly Enhanced Photoactivity of Anatase TiO2 Nanocrystals by Controlled Hydrogenation-Induced Surface Defects. ACS Catal. 2013, 3, 24792486. [38] Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R. C.; Wang, C. C.; Zhang, J. Z.; Li, Y. Hydrogen-Treated TiO2 Nanowire Arrays for Photoelectrochemical Water Splitting. Nano Lett. 2011, 11, 3026-3033. [39] Hoang, S.; Berglund, S. P.; Hahn, N. T.; Bard, A. J.; Mullins, C. B. Enhancing Visible Light

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry Letters

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

Page 26 of 33

Photo-oxidation of Water with TiO2 Nanowire Arrays via Cotreatment with H2 and NH3: Synergistic Effects between Ti3+ and N. J. Am. Chem. Soc. 2012, 134, 3659-3662. [40] Tian, Z. L.; Cui, H. L.; Zhu, G. L.; Zhao, W. L.; Xu, J. J.; Shao, F.; He, J. Q.; Huang, F. Q. Hydrogen plasma reduced black TiO2-B nanowires for enhanced photoelectrochemical watersplitting. J. Power. Sources. 2016, 325, 697-705. [41] Xu, Y. F.; Zhang, C.; Zhang, L.; Zhang, X.; Yao, H. L.; Shi, J. L. Pd-catalyzed instant hydrogenation of TiO2 with enhanced photocatalytic performance. Energy Environ. Sci. 2016, 9, 2410-2417. [42] Kang, Q.; Cao, J. Y.; Zhang, Y. J.; Liu, L. Q.; Xu, H.; Ye, J. H. Reduced TiO2 nanotube arrays for photoelectrochemical water splitting. J. Mater. Chem. A. 2013, 1, 5766-5774. [43] Cho, I. S.; Choi, J. M.; Zhang, K.; Kim, S. J.; Jeong, M. J.; Cai, L. L.; Park, T. H.; Zheng, X. L.; Park, J. H. Highly Efficient Solar Water Splitting from Transferred TiO2 Nanotube Arrays. Nano Lett. 2015, 15, 5709-5715. [44] Cho, I. S.; Logar, M.; Lee, C. H.; Cai, L. L.; Prinz, F. B.; Zheng, X. L. Rapid and Controllable Flame Reduction of TiO2 Nanowires for Enhanced Solar Water-Splitting. Nano Lett. 2013, 14, 24-31. [45] Mao, C. Y.; Zuo, F.; Hou, Y.; Bu, X. H.; Feng, P. Y. In Situ Preparation of a Ti3+ SelfDoped TiO2 Film with Enhanced Activity as Photoanode by N2H4 Reduction. Angew. Chem. Int. Ed. 2014, 53, 10485-10489. [46] Chen, H. N.; Wei, Z. H.; Yan, K. Y.; Bai, Y.; Yang, S. H. Unveiling Two ElectronTransport

Modes

in

Oxygen-Deficient

TiO2

Nanowires

and

Their

Influence

on

Photoelectrochemical Operation. J. Phys. Chem. Lett. 2014, 5, 2890-2896. [47] Yan, P. L.; Liu, G. J.; Ding, C. M.; Han, H. X.; Shi, J. Y.; Gan, Y.; Li, C.

ACS Paragon Plus Environment

26

Page 27 of 33

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

The Journal of Physical Chemistry Letters

Photoelectrochemical Water Splitting Promoted with a Disordered Surface Layer Created by Electrochemical Reduction. ACS Appl. Mater. Interfaces. 2015, 7, 3791-3796. [48] Pu, Y. C.; Ling, Y. C.; Chang, K. D.; Liu, C. M.; Zhang, J. Z.; Hsu, Y. J.; Li, Y. Surface Passivation of TiO2 Nanowires Using a Facile Precursor-Treatment Approach for Photoelectrochemical Water Oxidation. J. Phys. Chem. C. 2014, 118, 15086-15094. [49] Yang, Y.; Ling, Y.; Wang, G.; Liu, T.; Wang, F.; Zhai, T.; Tong, Y.; Li, Y. Photo-hole Induced Corrosion of Titanium Dioxide: Mechanism and Solutions. Nano Lett. 2015, 15, 70517057. [50] Zhang, K.; Ravishankar, S.; Ma, M.; Veerappan, G.; Bisquert, J.; Fabregat-Santiago, F.; Park, J. H. Overcoming Charge Collection Limitation at Solid/Liquid Interface by a Controllable Crystal Deficient Overlayer, Adv. Energy. Mater. DOI: 10.1002/aenm.201600923 [51] Zhang, K.; Wang, L. Y.; Kim, J. K.; Ma, M.; Veerappan, G.; Lee, C. L.; Kong, K. J.; Lee, H. Y.; Park, J. H. An order/disorder/water junction system for highly efficient co-catalyst-free photocatalytic hydrogen generation. Energy Environ. Sci. 2016, 9, 499-503. [52] Liu, N.; Schneider, C.; Freitag, D.; Hartmann, M.; Venkatesan, U.; Müller, J.; Spiecker, E.; Schmuki, P. Black TiO2 Nanotubes: Cocatalyst-Free Open-Circuit Hydrogen Generation, Nano Lett. 2014, 14, 3309-3313. [53] Liu, N.; Häublein, V.; Zhou, X. M.; Venkatesan, U.; Hartmann, M.; Mačković, M.; Nakajima, T.; Spiecker, E.; Osvet, A.; Frey, L.; Schmuki, P. “Black” TiO2 Nanotubes Formed by High-Energy Proton Implantation Show Noble-Metal-co-Catalyst Free Photocatalytic H2Evolution, Nano Lett. 2015, 15, 6815-6820. [54] Sun, Y. F.; Gao, S.; Xie, Y. Atomically-thick two-dimensional crystals: electronic structure regulation and energy device construction. Chem. Soc. Rev. 2014, 43, 530-546.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry Letters

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

Page 28 of 33

[55] Henderson, M. A. Structural Sensitivity in the Dissociation of Water on TiO2 Single-Crystal Surfaces. Langmuir. 1996, 12, 5093-5098. [56] Epling, W. S.; Peden, C. H. F.; Henderson, M. A.; Diebold, U. Evidence for oxygen adatoms on TiO2 (110) resulting from O2 dissociation at vacancy sites. Surf. Sci. 1998, 412-413, 333-343. [57] Wu, Q. P.; Huang, F.; Zhao, M.; Xu, J.; Zhou, J. C.; Wang, Y. S. Ultra-small yellow defective TiO2 nanoparticles for co-catalyst free photocatalytic hydrogen production, Nano Energy. 2016, 24, 63-71. [58] Li, R. G.; Weng, Y. X.; Zhou, X.; Wang, X. L.; Mi, Y.; Chong, R. F. Han, H. X.; Li, C. Achieving overall water splitting using titanium dioxide-based photocatalysts of different phases. Energy Environ. Sci. 2015, 8, 2377-2382. [59] Zhang, L.; Miller, B. K.; Crozier, P. A. Atomic Level In Situ Observation of Surface Amorphization in Anatase Nanocrystals During Light Irradiation in Water Vapor. Nano Lett. 2013, 13, 679-684. [60] Hu, S.; Shaner, M. R.; Beardslee, J. A.; Lichterman, M.; Brunschwig, B. S.; Lewis, N. S. Amorphous TiO2 coatings stabilize Si, GaAs, and GaP photoanodes for efficient water oxidation. Science. 2014, 344, 1005-1009. [61] Gu, J.; Yan, Y.; Young, J. L.; Xerxes Steirer, K.; Neale, N. R.; Turner, J. A. Water reduction by a p-GaInP2 photoelectrode stabilized by an amorphous TiO2 coating and a molecular cobalt catalyst. Nat. Mater. 2016, 15, 456-460. [62] Eisenberg, D.; Ahn, H. S.; Bard, A. J. Enhanced Photoelectrochemical Water Oxidation on Bismuth Vanadate by Electrodeposition of Amorphous Titanium Dioxide. J. Am. Chem. Soc. 2014, 136, 14011-14014.

ACS Paragon Plus Environment

28

Page 29 of 33

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

The Journal of Physical Chemistry Letters

[63] Yang, Y. Q.; Liu, G.; Irvine, J. T. S.; Cheng, H. M. Enhanced Photocatalytic H2 Production in Core–Shell Engineered Rutile TiO2. Adv. Mater. 2016, 28, 5850-5856. [64] Torquato, S.; Stillinger, F. H. Jammed hard-particle packings: From Kepler to Bernal and beyond. Rev. Mod. Phys. 2010, 82, 2633-2672. [65] Torquato, S.; Truskett, T. M; Debenedetti, P. G. Is random close packing of spheres well defined? Phys. Rev. Lett. 2000, 84, 2064. [66] Lu, H.; Zhao, B.; Pan, R.; Yao, J.; Qiu, J.; Luo, L.; Liu, Y. Safe and facile hydrogenation of commercial Degussa P25 at room temperature with enhanced photocatalytic activity. Rsc. Adv. 2014, 4, 1128-1132. [67] Zheng, Z.; Huang, B.; Meng, X.; Wang, J.; Wang, S.; Lou, Z.; Wang, Z.; Qin, X.; Zhang, X.; Dai, Y. Metallic zinc- assisted synthesis of Ti3+ self-doped TiO2 with tunable phase composition and visible-light photocatalytic activity. Chem. Commun. 2013, 49, 868-870. [68] Sinhamahapatra, A.; Jeon, J. P.; Yu, J. S. A new approach to prepare highly active and stable black titania for visible light-assisted hydrogen production. Energy Environ. Sci. 2015, 8, 3539-3544 [69] Zhu, G.; Yin, H.; Yang, C.; Cui, H.; Wang, Z.; Xu, J.; Lin, T.; Huang, F. Black Titania for Superior Photocatalytic Hydrogen Production and Photoelectrochemical Water Splitting. ChemCatChem. 2015, 7, 2614-2619. [70] Teng, F.; Li, M.; Gao, C.; Zhang, G.; Zhang, P.; Wang, Y.; Chen, L.; Xie, E. Preparation of black TiO2 by hydrogen plasma assisted chemical vapor deposition and its photocatalytic activity. Appl. Catal., B. 2014, 148, 339-343. [71] Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; Jiang, M. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry Letters

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

Page 30 of 33

aluminium-reduced black titania. Energy Environ. Sci. 2013, 6, 3007-3014. [72] Lü, X. J.; Chen, A. P.; Luo, Y. K.; Lu, P.; Dai, Y. M.; Enriquez, E.; Dowden, P.; Xu, H. W.; Kotula, P. G.; Azad, A. K.; Yarotski, D. A.; Prasankumar, R. P.; Taylor, A. J.; Thompson, J. D.; Jia, Q. X. Conducting Interface in Oxide Homojunction: Understanding of Superior Properties in Black TiO2, Nano lett. 2016, 16, 5751-5755. [73] Li, S.; Qiu, J.; Ling, M.; Peng, F.; Wood, B.; Zhang, S. Photoelectrochemical Characterization of Hydrogenated TiO2 Nanotubes as Photoanodes for Sensing Applications.. ACS Appl. Mater. Interfaces. 2013, 5, 11129-11135. [74] Danon, A.; Bhattacharyya, K.; Vijayan, B. K.; Lu, J.; Sauter, D. J.; Gray, K. A.; Stair, P. C.; Weitz, E. Effect of Reactor Materials on the Properties of Titanium Oxide Nanotubes. ACS Catal. 2011, 2, 45-49. [75] Li, J.; Liu, C. H.; Li, X.; Wang, Z. Q.; Shao, Y. C.; Wang, S. D.; Sun, X. L.; Pong, W. F.; Guo, J. H.; Sham, T. K. Unraveling the Origin of Visible Light Capture by Core-Shell TiO2 Nanotubes. Chem. Mater. 2016, 28, 4467-4475. [76] Walsh, F. C.; Wills, R. G. A. The continuing development of Magnéli phase titanium suboxides and Ebonex® electrodes. Electrochimica Acta. 2010, 55, 6342-6351. [77] Zhu, Y. M.; Liu, D. S.; Meng, M. H2 spillover enhanced hydrogenation capability of TiO2 used for photocatalytic splitting of water: a traditional phenomenon for new applications. Chem. Commun. 2014, 50, 6049-6051. [78] Nandasiri, M. I.; Shutthanandan, V.; Manandhar, S.; Schwarz, A. M.; Oxenford, L.; Kennedy, J. V.; Thevuthasan, S.; Henderson, M. A. Instability of hydrogenated TiO2. J. Phys. Chem. Lett. 2015, 6, 4627-4632. [79] Linsebigler, A. L.; Lu, G. Q.; Yates, J. T., Jr. Photocatalysis on TiO2 Surfaces: Principles,

ACS Paragon Plus Environment

30

Page 31 of 33

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

The Journal of Physical Chemistry Letters

Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735-758. [80] Thompson, T. L.; Yates, J. T., Jr. Surface Science Studies of the Photoactivation of TiO2New Photochemical Processes. Chem. Rev. 2006, 106, 4428−4453. [81] Angelis, F. D.; Valentin, C. D.; Fantacci, S.; Vittadini, A.; Selloni, A. Theoretical Studies on Anatase and Less Common TiO2 Phases: Bulk, Surfaces, and Nanomaterials. Chem. Rev. 2014, 114, 9708-9753. [82] Deskins, N. A.; Rousseau, R.; Dupuis, M. Distribution of Ti3+ Surface Sites in Reduced TiO2. J. Phys. Chem. C 2011, 115, 7562-7572. [83] Wang, Z. T.; Garcia, J. C.; Deskins, N. A.; Lyubinetsky, I. Ability of TiO2 (110) surface to be fully hydroxylated and fully reduced. Phys. Rev. B 2015, 92, 081402. [84] Setvin, M.; Aschauer, U.; Hulva, J.; Simschitz, T.; Daniel, B.; Schmid, M.; Selloni, A.; Diebold, U. Following the Reduction of Oxygen on TiO2 Anatase (101) Step by Step. J. Am. Chem. Soc., 2016, 138, 9565–9571. [85] Tao, H. B.; Fang, L.; Chen, J.; Yang, H. B.; Gao, J.; Miao, J.; Chen, S.; Liu, B. Identification of Surface Reactivity Descriptor for Transition Metal Identification of Surface Reactivity Descriptor for Transition Metal. J. Am. Chem. Soc. 2016, 138, 9978−9985. [86] S. Selcuk, A. Selloni, Facet-dependent trapping and dynamics of excess electrons at anatase TiO2 surfaces and aqueous interfaces. Nat. Mater. 2016, 15, 1107-1112. [87] Deskins, N. A.; Rousseau, R.; Dupuis, M. Localized Electronic States from Surface Hydroxyls and Polarons in TiO2 (110). J. Phys. Chem. C. 2009, 113, 14583-14586. [88] Liu, N.; Schneider, C.; Freitag, D.; Venkatesan, U.; Marthala, V. R. R.; Hartmann, M.; Winter, B.; Spiecker, E.; Osvet, A.; Zolnhofer, E. M.; Meyer, K.; Nakajima, T.; Zhou, X. M.; Schmuki, P. Hydrogenated Anatase: Strong Photocatalytic Dihydrogen Evolution without the

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry Letters

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

Page 32 of 33

Use of a Co-Catalyst. Angew. Chem. Int. Ed. 2014, 53, 14201-14205. [89] Xia, T.; Chen, X. B. Revealing the structural properties of hydrogenated black TiO2 nanocrystals. J. Mater. Chem. A, 2013, 1, 2983–2989. [90] Li, H.; Tsai, C.; Koh, A. L.; Cai, L.; Contryman, A. W.; Fragapane, A. H.; Zhao, J.; Han, H. S.; Manoharan, H. C.; Abild-Pedersen, F.; Nørskov, J. K.; Zheng, X. L. Activating and optimizing MoS2 basal planes for hydrogen evolution through the formation of strained sulphur vacancies. Nat. Mater. 2016, 15, 48–53. [91] Li, H.; Du, M. S.; Mleczko, M. J.; Koh, A. L.; Nishi, Y.; Pop, E.; Bard, A. J.; Zheng, X. L. Kinetic Study of Hydrogen Evolution Reaction over Strained MoS2 with Sulfur Vacancies Using Scanning Electrochemical Microscopy. J. Am. Chem. Soc., 2016, 138, 5123–5129. [92] Garcia, J. C.; Nolan, M.; Deskins, N. A. The nature of interfaces and charge trapping sites in photocatalytic mixed-phase TiO2 from first principles modeling. J. Chem. Phys. 2015, 142, 024708. [93] Xia, T.; Zhang, Chi.; Oyler, N. A.; Chen, X. B. Hydrogenated TiO2 Nanocrystals: A Novel Microwave Absorbing Material. Adv. Mater. 2013, 25, 6905–6910. [94] Shin, J. Y.; Joo, J. H.; Samuelis, D.; Maier, J. Oxygen-Deficient TiO2−δ Nanoparticles via Hydrogen Reduction for High Rate Capability Lithium Batteries. Chem. Mater. 2012, 24, 543551. [95] Lu, Z.; Yip, C. T.; Wang, L.; Huang, H.; Zhou, L. Hydrogenated TiO2 Nanotube Arrays as High-Rate Anodes for Lithium-Ion Microbatteries. ChemPlusChem. 2012, 77, 991-1000. [96] Shen, L.; Uchaker, E.; Zhang, X.; Cao, G. Hydrogenated Li4Ti5O12 Nanowire Arrays for High Rate Lithium Ion Batteries. Adv. Mater. 2012, 24, 6502-6506. [97] Myung, S. T.; Kikuchi, M.; Yoon, C. S.; Yashiro, H.; Kim, S. J.; Sun, Y. K.; Scrosati, B.

ACS Paragon Plus Environment

32

Page 33 of 33

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

The Journal of Physical Chemistry Letters

Black anatase titania enabling ultra high cycling rates for rechargeable lithium batteries. Energy Environ. Sci. 2013, 6, 2609-2614. [98] Lu, X.; Wang, G.; Zhai, T.; Yu, M.; Gan, J.; Tong, Y.; Li, Y. Hydrogenated TiO2 Nanotube Arrays for Supercapacitors. Nano Lett. 2012, 12, 1690-1696. [99] Li, H.; Chen, Z.; Tsang, C. K.; Li, Z.; Ran, X.; Lee, C.; Nie, B.; Zheng, L.; Hung, T.; Lu, J.; Pan, B.; Li, Y. Y. Electrochemical doping of anatase TiO2 in organic electrolytes for highperformance supercapacitors and photocatalysts. J. Mater. Chem. A. 2014, 2, 229-236. [100] Zhang, C.; Yu, H.; Li, Y.; Gao, Y.; Zhao, Y.; Song, W.; Shao, Z.; Yi, B. Supported noble metals on hydrogen-treated TiO2 nanotube arrays as highly ordered electrodes for fuel cells. ChemSusChem. 2013, 6, 659-666. [101] He, H. C.; Yang, K.; Wang, N.; Luo, F. F.; Chen, H. J. Hydrogenated TiO2 film for enhancing photovoltaic properties of solar cells and self-sensitized effect. J. Appl. Phys. 2013, 114, 213505. [102] Su, T.; Yang, Y. L.; Na, Y.; Fan, R. Q.; Li, L.; Wei, L. G.; Yang, B.; Cao, W. W. An Insight into the Role of Oxygen Vacancy in Hydrogenated TiO2 Nanocrystals in the Performance of Dye-Sensitized Solar Cells, ACS Appl. Mater. Interfaces. 2015, 7, 3754-3763. [103] Nolan, M.; Iwaszuk, A.; Lucid, A. K.; Carey, J. J.; Fronzi, M. Design of Novel Visible Light Active Photocatalyst Materials: Surface Modified TiO2. Adv. Mater. 2016, 28, 5425–5446. [104] Ade, H.; Stoll, H. Near-edge X-ray absorption fine-structure microscopy of organic and magnetic materials. Nat. Mater. 2009, 8, 281-290.

ACS Paragon Plus Environment

33