Bridging the g-C3N4 Interlayers for Enhanced ... - ACS Publications

Mar 3, 2016 - Technology and Business University, Chongqing 400067, China ... and Low Carbon Technology, Sichuan University, Chengdu 610065, China...
0 downloads 0 Views 2MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

Bridging the g-C3N4 interlayers for enhanced photocatalysis Ting Xiong, Wanglai Cen, Yuxin Zhang, and Fan Dong ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.5b02922 • Publication Date (Web): 03 Mar 2016 Downloaded from http://pubs.acs.org on March 7, 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.

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 26

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 Catalysis

1

Bridging the g-C3N4 interlayers for enhanced photocatalysis

2 3 4

Ting Xiong a,†, Wanglai Cen b,†, Yuxin Zhang c, Fan Dong a,*

5

a

6

Environment and Resources, Chongqing Technology and Business University, Chongqing

7

400067, China.

8 9 10 11

Chongqing Key Laboratory of Catalysis and Functional Organic Molecules, College of

b

Institute of New Energy and Low Carbon Technology, Sichuan University, Chengdu 610065, China.

12

c

13

Science of Micro/Nano-Devices and System Technology, Chongqing University, Chongqing

14

400044, China.

College of Materials Science and Engineering, National Key Laboratory of Fundamental

1 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2

ABSTRACT: Graphitic carbon nitride (g-C3N4) has been widely investigated and applied in

3

photocatalysis and catalysis, but its performance is still unsatisfactory. Here, we demonstrated

4

that K-doped g-C3N4 with unique electronic structure possessed highly enhanced visible light

5

photocatalytic performance for NO removal, which was superior to Na-doped g-C3N4. DFT

6

calculations revealed that K or Na doping can narrow the bandgap of g-C3N4. K atoms,

7

intercalated into the g-C3N4 interlayer via bridging the layers, could decrease the electronic

8

localization and extend the π conjugated system, while Na atoms tended to be doped into the

9

CN planes and increased the in-planar electron density. Based on theoretical calculation

10

results, we synthesized K-doped g-C3N4 and Na-doped g-C3N4 by a facile thermal

11

polymerization method. Consistent with the theoretical prediction, it was found that K was

12

intercalated into space between the g-C3N4 layers. The K-intercalated g-C3N4 sample showed

13

increased visible light absorption, efficient separation of charge carriers and strong oxidation

14

capability, benefiting from the narrowed band gap, extended π conjugated systems and

15

positive-shifted valence band position, respectively. Despite that the Na-doped g-C3N4

16

exhibited narrowed bandgap, the high recombination rate of carriers resulted in the reduced

17

photocatalytic performance. Our discovery provides a promising route to manipulate the

18

photocatalytic activity simply by introducing K atoms in the interlayer and gains a deep

19

understanding of doping chemistry with congeners. The present work could provide new

20

insights into the mechanistic understanding and the design of electronically optimized layered

21

photocatalysts for enhanced solar energy conversion.

22 23

KEYWORDS: K-intercalated g-C3N4; bridging K atoms; charge redistribution and transfer;

24

thermal polymerization; visible light photocatalysis.

2 ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26

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 Catalysis

1

1. INTRODUCTION

2

Recently, beyond conventional semiconductors, two-dimensional (2D) materials are of

3

particular interests in consideration of their fantastic, unusual properties and potential

4

applications in various fields.1-3 Typically, graphene, graphitic carbon nitride, hexagonal

5

boron nitride and transition metal dichalcogenides with remarkable properties have been

6

successfully applied in optical and electronic devices, energy generation and storage,

7

environmental remediation, hybrid materials as well as chemical sensors.4-8

8

Among the various available 2D layered compounds, g-C3N4, consisting of tris-triazine

9

units connected with planar amino groups in each layer and weak van der Waals force

10

between layers (Figure 1), has triggered extensive investigation due to the suitable bandgap,

11

low cost, ease of preparation, good stability, and environmental friendly feature, thus leading

12

to multifunctional application for photocatalytic degradation of pollutants, photocatalytic

13

hydrogen generation, carbon dioxide reduction as well as supercapacitors.9-14 Nevertheless,

14

the pure g-C3N4 has been restricted by the low photocatalysis efficiency mainly for its

15

marginal visible light absorption and fast charge recombination.15,16 In order to improve the

16

photocatalytic performance of g-C3N4, various modification strategies have been

17

employed,17-25 and doping is known to considerably broaden the light responsive range and

18

enhance the charge separation of semiconductors. Doping of g-C3N4 with S, F, C or B via

19

substituting for lattice atoms has been applied to modify its texture and electronic structure

20

for improving the photocatalytic performance.26-29 Also, transitional metals (Fe3+, Mn3+, Co3+,

21

Ni3+, and Cu2+) have been incorporated into the framework (nitrogen pores) of g-C3N4 to

22

enhance the performance in benzene hydroxylation and styrene epoxidation reaction.30-32 As

23

can be seen, the externally doped atoms either substitute for the lattice atoms or exist in the

24

in-planar caves of g-C3N4.

25

For layered materials, intercalation provides the materials with promising properties. For

26

example,

calcium-intercalated

bilayer

graphene

C6CaC6

27

superconductivity since the interlayer electrons participate in the transport properties.33 FeCl3

28

intercalated few-layer graphene has a square resistance lower than ITO.34 Also, it is found

29

that Pt-intercalated layered KCa2Nb3O10 shows efficient photocatalytic activity for water

30

splitting.35 Given that g-C3N4 has a layered structure with interlayer galleries that could allow 3 ACS Paragon Plus Environment

on

silicon

possesses

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

doping of heteroatom within the interlayer, it is a feasible route to synthesize intercalated

2

g-C3N4 compound to achieve enhanced photocatalytic performance. Yan et al. prepared

3

intercalated carbon nitride photocatalysts for hydrogen production, which is unstable in

4

aqueous solution.36 Hence, it still remains a challenge to fabricate stable and superior

5

intercalated g-C3N4 compound.

6

Recently, several literatures reported that the introduction of K atoms can improve the

7

photocatalytic performance of g-C3N4.37,38 However, the position of the introduced K atoms

8

is not clear and the essential evidence for the improved photocatalytic performance has not

9

been revealed clearly. Furthermore, whether other congener alkaline elements, for instance

10

the Na atoms with similar electronic structure, possess similar promotion effects, are of great

11

interest to be uncovered. Herein, we performed DFT calculations to investigate the electronic

12

and band structures of K/Na-doped g-C3N4. It is amazing to find that the bandgap of g-C3N4

13

could be narrowed either by K or Na doping, but they exerted different influence on the

14

electronic structure and photocatalytic activity of g-C3N4. The K atoms tended to exist in the

15

g-C3N4 interlayer and extended the π conjugated systems, while Na atoms were doped into

16

the conjugated plane and increased the electron density in the CN layers. The structural

17

differences may lead to the different photocatalytic performance. Aimed to this point, we

18

prepared K/Na-doped g-C3N4 by a simple thermal polymerization method. As expected, the

19

K-intercalated g-C3N4 exhibited high photocatalytic activity toward the removal of NO.

20

However, Na-doped g-C3N4 showed low photocatalytic activity. A series of characterizations

21

indicated that the K-intercalated g-C3N4 possessed narrowed band gap and strong VB holes

22

oxidation. The intercalated K atoms benefited the transfer and separation of charge carriers.

23

However, it was not the way for Na, although the Na doped g-C3N4 also exhibited narrowed

24

band gap. The present work could provide new insights into the effects of alkali metal doping

25

on g-C3N4 as well as the design of intercalated photocatalysts with highly efficient

26

visible-light-driven activity for air purification.

27

2. EXPERIMENTAL

28

2.1 Materials and synthesis

29

All chemicals used in this study were analytical grade and were used without further

30

treatment. In a typical synthesis procedure, 10 g of thiourea, and a known amount (3 wt %) 4 ACS Paragon Plus Environment

Page 4 of 26

Page 5 of 26

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 Catalysis

1

(relative to the experimentally obtained g-C3N4) of X (X = Na, K) coming from X (X = Na,

2

K)Br were dissolved in 30 mL water in an alumina crucible. The obtained solution was then

3

dried at 80 °C overnight to get the solid precursors. The solid composite precursors were

4

placed in a semi-closed alumina crucible with a cover. The crucible were heated to 550 °C at

5

a heating rate of 15 °C·min-1 in a muffle furnace and maintained for 2 h. After the thermal

6

treatment, the crucible was cooled down to room temperature in the muffle furnace. The

7

resulted samples were collected for further use. The pristine g-C3N4 was prepared without

8

adding XBr and labeled as CN. The Na-doped g-C3N4 was labeled as CN-Na3. The K-doped

9

g-C3N4 with different weight ratio of K to g-C3N4 (1, 3, 5 and 10 wt %) was labeled as

10

CN-K1, CN-K3, CN-K5 and CN-K10, respectively.

11

2.2 Characterization

12

The crystal phases of the sample were analyzed by X-ray diffraction (XRD) with Cu Kα

13

radiation (model D/max RA, Rigaku Co., Japan). Scanning electron microscopy (SEM;

14

model JSM-6490, JEOL, Japan) was used to characterize the morphology of the obtained

15

products. The morphology and structure of the samples were examined by transmission

16

electron microscopy (TEM; JEM-2010, Japan). X-ray photoelectron spectroscopy (XPS) with

17

Al Kα X-rays (Thermo ESCALAB 250, USA) was used to investigate the surface properties.

18

The UV-vis diffuse-reflectance spectrometry (DRS) spectra were obtained for the dry-pressed

19

disk samples using a Scan UV-vis spectrophotometer (TU-1901, China) equipped with an

20

integrating sphere assembly, using 100% BaSO4 as the reflectance sample. Nitrogen

21

adsorption-desorption isotherms were obtained on a nitrogen adsorption apparatus (ASAP

22

2020, USA) with all samples degassed at 150 °C for 12 h prior to measurements. The

23

photoluminescence spectra were measured with a fluorescence spectrophotometer (F-7000,

24

Japan) using a Xe lamp as excitation source with optical filters. Steady and time-resolved

25

fluorescence emission spectra were recorded at room temperature with a fluorescence

26

spectrophotometer

27

thermogravimetric-differential scanning calorimetry analysis (TG-DSC: NETZSCHSTA 409

28

PC/PG), 20 mg dry sample was sealed in an Al2O3 crucible with a lid and scanned at a rate of

29

20 °C·min−1. The photocurrent response and electrochemical impedance spectra

30

measurements were conducted in a three electrode system on a CH 660D electrochemical

31

work station, Platinum wire was used as the counter electrode. Sturated calomel electrodes

32

were used as the reference electrodes. The as-prepared samples film electrodes on ITO served 5

(Edinburgh

Instruments,

FLSP-920).

ACS Paragon Plus Environment

To

perform

the

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

as the working electrode. All the potentials quoted here are with respect to saturated calomel

2

electrode. For photoelectrochemical test, the working electrode was irradiated from the

3

as-prepared sample films side under a 300 W Xe lamp (PerfectLight Co. Ltd,

4

PLS-SXE-300UV). Incident visible-light was obtained by utilizing a 420 nm cutoff filter. The

5

photocurrent-time dependence of as-prepared sample films at open circuit potential (OCV)

6

was measured in 0.5 M Na2SO4 under chopped illumination with 60s light on/off cycles.

7

2.3 DFT Calculations

8

DFT calculations were carried out using the “Vienna ab initio simulation package”

9

(VASP5.3).39 The Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional was used

10

within the spin polarized generalized gradient approximation (GGA).40 A plane-wave basis

11

set was employed within the framework of the projector augmented-wave (PAW) method.41

12

In order to get accurate results, the cut-off was set to 450 eV. A Gaussian smearing was used

13

with a smearing width of 0.2 eV. Geometry relaxations were carried out until the residual

14

forces on each ion were smaller than 0.02 eV/Å. We firstly relaxed a 1×1×2 supercell of bulk

15

g-C3N4. Then, a single K or Na atom was introduced in interstitial or substitutional doping

16

ways. The most stable relaxed configurations are the in plane doping pattern for Na, and the

17

interlayer bridging pattern for K according to our testing calculations (see the supporting

18

information). The most stable relaxed configuration was denoted as CN-K2 (CN-Na2). It

19

results in a concentration of 3 wt% Na for CN-Na2 or 5 wt% K for CN-K2. The electronic

20

structures were calculated based on the fully relaxed lattice parameters and ionic positions. In

21

all calculations, k-points were sampled in 5×5×3 Monkhorst-Pack grid. The van der Waals

22

(vdW) correction was included by using the DFT-D2 approach.42

23

2.4 Evaluation of visible light photocatalytic activity

24

The photocatalytic activity was investigated by removal of NO at ppb levels in a

25

continuous flow reactor at ambient temperature. The volume of the rectangular reactor, made

26

of polymeric glass and covered with Saint-Glass, was 4.5 L (30 cm × 15 cm × 10 cm). A 150

27

W commercial tungsten halogen lamp was vertically placed outside the reactor. A UV cutoff

28

filter (420 nm) was adopted to remove UV light in the light beam. For each photocatalytic

29

activity test, 0.20 g of the as-prepared sample was dispersed in distilled water (50 mL) in a

30

beaker via ultrasonic treatment for 10 min and then coated onto two glass dishes (12.0 cm in 6 ACS Paragon Plus Environment

Page 6 of 26

Page 7 of 26

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 Catalysis

1

diameter). The coated dish was pretreated at 60 °C to remove water in the suspension and

2

then cooled to room temperature before photocatalytic test.

3

The NO gas was acquired from a compressed gas cylinder at a concentration of 100 ppm of

4

NO (N2 balance). The initial concentration of NO was diluted to about 600 ppb by the air

5

stream. The desired relative humidity level of the NO flow was controlled at 50% by passing

6

the zero air streams through a humidification chamber. The gas streams were premixed

7

completely by a gas blender, and the flow rates of the air stream and NO were controlled at

8

2.4 L·min-1 and 15 mL·min-1 by a mass flow controller, respectively. After the

9

adsorption-desorption equilibrium was achieved, the lamp was turned on. The concentration

10

of NO was continuously measured by a chemiluminescence NO analyzer (Thermo

11

Environmental Instruments Inc., 42i-TL), which monitors NO with a sampling rate of 1.0

12

L·min-1. The removal ratio (η) of NO was calculated as η (%) = (1-C/C0) ×100%, where C

13

and C0 are concentrations of NO in the outlet steam and the feeding stream, respectively.

14

3. RESULTS AND DISCUSSION

15

3.1 DFT calculation and electronic structure

16

7 ACS Paragon Plus Environment

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 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 26

1 2

Figure 1. The calculated crystal structures of CN (a), CN-K2 (b) and CN-Na2 (c). (d) and (e)

3

are the topviews of the doped layer in (b) and (c), correspondingly. Selected distances are

4

marked in pm.

5

DFT calculation was first utilized to predict the crystal structure of g-C3N4 without and

6

with K or Na atoms doping. The optimal g-C3N4 crystal structure is shown in Figure 1a.

7

Related lattice parameters were collected in Table 1. The in-planar distance of the nitride

8

pores at the center of adjacent heptazine units and the interlayer distance of the g-C3N4 are

9

determined to be ca. 7.11 and 3.14 Å, respectively, in agreement with the other theoretical

10

results.43,44 Based on the optimal crystal structure of g-C3N4, we tested both the in cave

11

doping and interlayer bridging sites for Na and K. Related results have been provided in the

12

supporting information (Table S1-S2 and Figure S1-S4). The most stable relaxed

13

configuration is the in cave doping pattern for Na, and the interlayer bridging pattern for K.

14

The optimized crystal structures of Na doped g-C3N4 (CN-Na2) and K intercalated g-C3N4

15

(CN-K2) are shown in Figure 1b and 1d, Figure 1c and 1e, respectively. Obviously, after

16

structural optimization, the Na atoms are found to locate in the π-conjugated planes, while K

17

atoms are present in the g-C3N4 interlayer. Moreover, Na doping almost does not change the

18

crystal structure, but the crystal cell becomes slim after K doping. Such different influence

19

induced by Na and K doping mainly originates from the size of the two atoms. Na atoms with

20

small radius of 1.86 Å can enter the cave of the planes, 45 and then exert little influence on the

21

crystal structure. K atoms with relatively large radius of 2.32 Å are inclined to intercalate into

22

g-C3N4 interlayer, which are expected to form a bridge for charge transfer.45

23

Table 1. Comparison of the calculated crystal structures over CN, CN-Na2 and CN-K2. a (Å)

b (Å)

c (Å)

α (o)

β (o)

γ (o)

CN

7.11

7.11

12.26

90.00

90.00

120.00

CN-Na2

6.98

7.00

12.56

90.00

89.94

120.14

CN-K2

6.98

7.02

13.18

90.00

91.90

120.24

Parameters

8 ACS Paragon Plus Environment

Page 9 of 26

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 Catalysis

CN43

7.16

6.96

12.35

90.00

90.00

120.00

CN44

7.13

6.92

12.35

90.00

90.00

120.00

1 2

Figure 2. The charge difference distribution of (a) CN-K2 and (b) CN-Na2, and electronic

3

location function (ELF) analysis of (c) CN-K2 and (d) CN-Na2. In (a) and (b), the yellow

4

region denotes charge depletion while blue region denotes accumulation. The isosurface is

5

0.02 e/Å3. Refer to Figure S5 in the supporting information for full views.

6

Then we analyzed the differences in spatial charge distribution and ELF of CN-K2 and

7

CN-Na2 as shown in Figure 2. Although Na and K atoms are congeners, they exert distinct

8

influence on the charge distribution property. All C-N bonds are sp2 orbital hybridization and

9

the lone electrons in the pz orbital form a big π-bond in g-C3N4.9 In the host g-C3N4, valence

10

electrons are preferable to locate at N atoms for its stronger eletrophilicity than C atom,

11

which results in the formation of N anion. When K and Na are introduced, the outermost

12

electrons are transferred to g-C3N4 and the Na/K presents as cation. Subsequently, there is

13

static coulomb interaction between Na/K and N nearest to them. K atoms lead to the drastic

14

rearrangement of the g-C3N4 layer. The sp2 orbital planes lose electrons and the pz orbital

15

planes obtain electrons in CN-K2 (Figure 2a and Figure S5a). The charge of the upper layer 9 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

N (C) atoms and the lower layer C (N) atoms close to the interbedded K atoms increases,

2

forming a funnel-like three dimensional structure. ELF is defined as a dimensionless quantity

3

and takes values in the range between 0 and 1, where ELF = 1 corresponds to the perfect

4

localization and ELF = 0.5 is for the uniform electron gas. It can be used as an indicator of

5

covalent bond between two atoms.46 The ELF of CN-K2 is displayed in Figure 2c, the

6

minimum value of ELF between K-N and K-C are quite small (420 nm, [Na2SO4] = 0.5 M).

4

Figure S8 shows the XRD pattern of CN, CN-Na3 and CN-K3. Notably, compared with

5

the pure g-C3N4, the peaks position for Na-doped g-C3N4 keep unchanged, whereas the peaks

6

position experience a shift for K-doped g-C3N4. This experimental result confirms the

7

theoretical result that K is intercalated g-C3N4 and changes the layer distance but Na is doped

8

into g-C3N4 and has no obvious influence on the layer distance. The photocurrent response

9

and electrochemical impedance spectra measurements were carried out to distinguish the

10

variations of photoelectric response after Na or K doping. In contrast to CN, CN-K3 exhibits

11

an improved photocurrent, but CN-Na3 shows a decreased photocurrent under visible light

12

irradiation (Figure 5a). The increased photocurrent suggests more efficient separation of

13

photogenerated electron/hole pairs.47 In addition, EIS Nyquist analysis were conducted as

14

shown in Figure 5b. The diameter of the arc radius on the Nyquist plots of the K-intercalated

15

g-C3N4 is smaller than that of the bare g-C3N4, whose diameter is smaller than that of

16

Na-doped g-C3N4. Figure S9 displays that the PL intensity of the three samples follows the

17

order of and CN-K3 < CN < CN-Na3. It turns out that introducing K atoms into g-C3N4 can

18

facilitate the separation and transfer efficiency of photogenerated carriers, whereas

19

introducing Na atoms plays an opposite role.

13 ACS Paragon Plus Environment

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 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 26

1 2

Figure 6. The ns-level time-resolved fluorescence spectra monitored under 420 nm excitation

3

at room temperature for CN, CN-Na3 and CN-K3.

4

The ns-level time-resolved fluorescence decay spectra were further employed to

5

investigate the charge transfer dynamics over CN, CN-Na3 and CN-K3 (Figure 6). The

6

curves can be fitted well to a biexponential decay function, with all the fitting parameters

7

summarized in Table 2. The short lifetime (τ1) is 10.5, 6.3 and 14.0 ns, and the long lifetime

8

(τ2) is 85.6, 80.6 and 90.3 ns for CN, CN-Na3 and CN-K3, respectively. Obviously, in

9

contrast to CN, the lifetime of charge carriers is prolonged for CN-K3 while that for CN-Na3

10

is shortened, elucidating that K doping can enhance the charge transfer and separation of

11

g-C3N4 while the Na doping could not do so. The prolonged lifetime of charge carriers can

12

increase their probability to participate in photocatalytic reactions before recombination.48

13

These results combined with the analysis revealed by the photocurrent responses, EIS spectra

14

and PL spectra confirm the conclusion that the fast transfer and separation of charge carriers

15

are responsible for the high photocatalytic activity of CN-K3 while the high recombination of

16

carriers results in low activity for CN-Na3.

17

Table 2. Kinetic parameters of the fitting decay parameters of CN, CN-Na3 and CN-K3. Relative Samples

Parameters

Life time (ns)

χ2

Percentage (%) τ1

10.5

54.3

τ2

85.6

45.7

τ1

6.3

70.9

CN CN-Na3

1.041

14 ACS Paragon Plus Environment

1.056

Page 15 of 26

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 Catalysis

τ2

80.6

29.1

τ1

14.0

49.5

τ2

90.3

50.5

CN-K3

1

1.065

3.4 Phase Structure and Chemical Composition

2 3 4

Figure 7. The XRD of g-C3N4 and the K-intercalated g-C3N4 samples with different K doping content (a) and the enlarged profile of the (002) diffraction region (b).

5

As the K-intercalation could significantly enhance the photocatalytic activity of g-C3N4,

6

we prepared a series of K-intercalated g-C3N4 photocatalysts by tuning the amount of K to

7

further optimize the photocatalyst. Figure 7a depicts the X-ray diffraction (XRD) patterns of

8

the samples. Two peaks at ca. 13.1° (indexed as (100) diffraction planes associated with the

9

in-plane repeated unites) and 27.7° (corresponding to the (002) peak of the periodic graphitic

10

stacking of the conjugated aromatic system) can be observed in the g-C3N4 and

11

K-intercalated g-C3N4 samples, reflecting the existence of graphitic-like layer structures.9

12

Also noted is the downshift of the (002) diffraction peaks in K-intercalated g-C3N4 compared

13

with pure g-C3N4 with the increase of K doping content (Figure 7b). This could be attributed

14

to the incorporation of K atoms between the CN layers. K atoms existing in g-C3N4 interlayer

15

are bonded with the adjacent C or N atoms, which will then induce the expansion of the

16

crystal lattice for g-C3N4. This is confirmed by the theoretic result that obvious change of

17

crystal structure is observed for K-intercalated g-C3N4.

15 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

2 3

Figure 8. XPS spectra of CN and CN-K5, (a) C 1s, (b) N 1s, (c) K 2p, (d) Br 3d.

4

The compositions and chemical states of CN and CN-K5 were further tested by X-ray

5

photoelectron spectroscopy (XPS) for comparison. In Figure 8a, the two bonding states of

6

carbon species are evidenced with the C 1s binding energies of around 284.8 and 288.2 eV.

7

The former is assigned as the surface adventitious carbon or sp2 C–C bonds, and the latter is

8

attributed to the sp2-bonded carbon in N-containing aromatic rings.49 The high resolution N1s

9

XPS spectra of the two samples are shown in Figure 8b. The predominant peaks at a binding

10

energy of ca. 398.5 eV can be ascribed to the sp2 hybridized nitrogen involved in triazine

11

rings (C–N=C). Whereas the other two peaks with binding energies of 401.1 and 404.1 eV

12

are assigned to the N-H groups and the charging effects about the π-excitations, separately.50

13

Furthermore, the K element with binding energy located at 293 and 295.7 eV corresponding

14

to the K 2p3/2 and K2p1/2 peaks (Figure 8c) is observed in CN-K5, suggesting the presence of

15

the K species in the form of K-N and K-C bonds.51,52 The atomic concentration of the doped

16

K is determined to be 4.3% by XPS, close to the theoretic doped amount. The Br element is 16 ACS Paragon Plus Environment

Page 16 of 26

Page 17 of 26

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 Catalysis

1

not doped into g-C3N4 as it is not detected in CN-K5 (Figure 8d). Notably, there are obvious

2

chemical shifts in the predominant N 1s and C1s peaks in relative to the pure g-C3N4. The

3

observed shifts for N 1s and C 1s binding energy further indicate the strong effects existing

4

between K and C/N atoms, which agree well with the theoretic analysis. These results

5

coupled with XRD analysis demonstrate the successful synthesis of K-intercalated g-C3N4.

6

3.5 Thermo stability

7 8

Figure 9. TG (a) and DSC (b) of the CN and CN-K3 samples.

9

In order to understand the thermal properties of CN and CN-K3, the thermogravimetric

10

(TG) and differential scanning calorimetry (DSC) analysis were carried out. The detected

11

range of temperature was from room temperature to 1000 °C at a heating rate of 20 °C·min-1.

12

In the case of TG, the weight loss of less than 10% at temperatures of up to 200 °C is

13

associated with the removal of physically adsorbed water. While a drastic weight loss is

14

observed in the range of 530–780 °C, which can be attributed to the loss in

15

tri-s-triazine-based units or other advanced condensates (Figure 9a).53 Correspondingly, the

16

DSC thermogram also witnesses the endothermic and exothermic stages in the same ranges

17

as those obtained in the TG thermogram (Figure 9b). It is noted that the major endothermic

18

peak at 743 °C experiences a downshift for CN-K3 in comparison with the pure g-C3N4,

19

implying that the stability of g-C3N4 decreases slightly after K doping. This further

20

demonstrates that K atoms are incorporated into the interlayer and create enlarged interlayer

21

distance, and consequently decrease the stability. 17 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Page 18 of 26

3.6 Morphological Structure

(b)

(a)

2

(d)

(c)

3

N

C

(e)

K

4

Figure 10. SEM and TEM images of CN (a) and CN-K5 (b), FESEM-EDX elemental

5

mapping of CN-K5 (c, d and e).

6

The typical SEM and TEM images of CN and CN-K5 are shown in Figure 10. Figure 10a

7

shows that g-C3N4 sample is composed of irregularly curved layers. These layers are packed

8

in a random way, resulting in porous structure. Figure 10b illustrates that the CN-K5, similar

9

to g-C3N4, is also composed of curved layers, which implies that introduction of K atoms into

10

the crystal structure does not influence the morphology. Besides, elemental mapping based

11

CN-K5 was performed. It reveals that there appears a distribution of C, N, and K across the

12

bulk structure (Figure 10c, 10d and 10e). The K signal is detected in the entire structure,

13

verifying the homogenous dispersion of K element in g-C3N4.

14

3.7 Light Absorption and Band Structure

18 ACS Paragon Plus Environment

Page 19 of 26

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 Catalysis

1

(d)

2 3

Figure 11. UV-vis spectra (a) and the estimated band gaps (b) of g-C3N4 and the

4

K-intercalated g-C3N4 samples, VB XPS (c) and schematic illustration of the band gap

5

structure of CN and CN-K5 (d).

6

The light absorption property of these samples was investigated as shown in Figure 11a. In

7

contrast to the pure g-C3N4, the K-intercalated g-C3N4 samples show enhanced visible

8

absorption, and the absorption edges of K-intercalated g-C3N4 samples undergo red shifts.

9

Evidently, CN-K5 exhibits the strongest visible light absorption. The band gaps of the

10

as-synthesized CN, CN-K1, CN-K3, CN-K5 and CN-K10 (Eg) samples estimated from the

11

intercept of the tangents to the plots of (αhν)1/2 νs. photoenergy (Figure 11b) are 2.45, 2.34,

12

2.19, 2.15 and 2.15 eV, respectively, confirming that the bandgap can be narrowed by

13

introducing K atoms into g-C3N4 interlayers.54 Estimated from Figure 11c, the VB edges of

14

g-C3N4 and CN-K5 are estimated to be 1.58 and 1.86, respectively.55,56 According to the

15

bandgap and the VB position of the two samples, we can draw the bandgap structures as 19 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

displayed in Figure 11d. It is obviously that the CB and VB position down-shift over CN-K5.

2

These facts that K-intercalated g-C3N4 samples possess narrowed band gap and enhanced

3

oxidization ability of VB holes relative to g-C3N4 are in line with the DOS results. These

4

experimentally determine that band structures are well consistent with the DFT calculation

5

results (Figure 3).

6

3.8 Carriers separation and transfer

7 8

Figure 12. PL spectra of the as-prepared samples.

9

To further understand the transfer and recombination processes of photoexcited charge

10

carriers in these K-intercalated g-C3N4 samples, PL spectra were measured. As displayed in

11

Figure 12, the K-intercalated g-C3N4 samples give similar PL spectra to pristine g-C3N4. But

12

the PL emission intensity over K-intercalated C3N4 is decreased with respect to g-C3N4, and

13

the intensity decreases with the increasing K content. Namely, the recombination of the

14

photoexcited charge carriers is effectively inhibited by introduction of K atoms into the

15

g-C3N4 interlayer. Additionally, the BET specific surface areas (SBET) of CN, CN-K1, CN-K3,

16

CN-K5 and CN-K10 products are determined to be 27, 18, 14, 11 and 8 m2·g−1, respectively.

17

It can be seen that the K-intercalated g-C3N4 samples show slightly lowered SBET in contrast

18

to the pure g-C3N4.

19

Table 3. The SBET, Band gap, interlayer distance, NO removal ratio and k values of all 20 ACS Paragon Plus Environment

Page 20 of 26

Page 21 of 26

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

1

2

ACS Catalysis

K-intercalated g-C3N4 samples Sample

SBET /m2·g

Band gap /eV

Interlayer distance /nm

η(NO) /%

k value /min−1

CN

27

2.45

0.322

22.4

0.0939

CN-K1

18

2.34

0.323

33.5

0.1195

CN-K3

14

2.19

0.324

34.9

0.1138

CN-K5

11

2.15

0.325

36.8

0.1061

CN-K10

8

2.15

0.326

27.0

0.1082

3.9 Photocatalytic activity and stability

3 4

Figure 13. Comparison of the visible photocatalytic performance over the g-C3N4 and

5

K-intercalated g-C3N4 (a) and the cycling test of CN-K5 (b).

6

The photocatalytic performances of g-C3N4 and K-intercalated g-C3N4 samples have been

7

tested toward the removal of NO under visible light irradiation. Previous researches reveal

8

that NO could not be photolyzed without photocatalysts under visible light irradiation.57 As

9

can be seen from Figure 13a, for all samples, the NO removal ratios reach maximum in 5 min

10

and then decrease and finally reach a stable value with the following irradiation time. The

11

reduced activity may be caused by the accumulated reaction intermediates and products

12

generated on the surface during photocatalytic reaction. With the presence of pure g-C3N4,

13

NO removal ratio of 16% can be achieved after 30 min irradiation. Introducing K atoms into 21 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

g-C3N4 leads to enhanced visible photocatalytic performance. It is found that all the

2

K-intercalated g-C3N4 samples display boosted activity and high reaction rate constants k

3

(Figure S10), which are superior to g-C3N4 and the physically mixed g-C3N4 and KBr

4

(summarized in Table 3). CN-K5 possesses the highest NO removal ratio of 37%, indicating

5

that there is an optimal K doping content to improve the activity of g-C3N4. The high visible

6

photocatalytic activity is ascribed to the intercalated K atoms. For one hand, K atoms

7

introduced into the g-C3N4 interlayer not only narrow the band gap and enhance visible light

8

absorption, but also downshift the VB position to improve the holes’ oxidization capability.

9

On the other hand, K atoms located at the g-C3N4 interlayer, which can function as delivery

10

channels for charge carries and then prolong the lifetime of photogenerated holes and

11

electrons. Based on previous study, ·O2−, h+ and ·OH are found to be the active species

12

involved with NO removal reactions.12 However, excess K doping may create nonradiative

13

recombination sites and decrease the photocatalytic activity.58,59 In addition, the catalytic

14

stability of CN-K5 was evaluated by six consecutive tests (Figure 13b). After two consecutive

15

tests, CN-K5 shows slightly decreased photocatalytic performance. Later, the NO removal

16

ratio keeps constant without noticeable deactivation, and exceeds the photocatalytic

17

performance of g-C3N4.

18

On the basis of all the results, we could gain a deeper understanding on the alkaline metals

19

doping and their differences. K and Na doping can narrow the band gap of g-C3N4. Meantime,

20

the band edges experience downshifts. Importantly, the congener alkaline metals doping exert

21

different effects on the electronic structures. For example, K with large size tends to be doped

22

into the interlayer while Na is doped into the CN layers. The intercalated K bonded with

23

atoms at the adjacent layers could form the delivery channels, increase the electronic

24

delocalization and extend the π conjugated systems, which will contribute to the transfer and

25

separation of charge carriers. Nevertheless, Na doping increases the in-planar electron density

26

and then increases the recombination of the charge carriers. In addition, K possesses stronger

27

metallicity than Na atoms. Na atoms combine with in-planar N atoms with ionic bonds,

28

whereas the interbedded K atoms are ionically and covalently combined with the N and C

29

atoms at the adjacent layers. This is the first report to enhance the charge separation via

30

bridging the layers with alkaline metal. This work is essential for better understanding the 22 ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26

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 Catalysis

1

congener element doping-activity relationships as well as for the mechanistic understanding

2

of photocatalysis and rational design of potent photocatalysts.

3 4

4. CONCLUSION

5

We revealed the effects of K and Na doping on the electronic structure and photocatalytic

6

activity of g-C3N4 with a combined experimental and theoretical approach. The band gap of

7

g-C3N4 was narrowed by introducing K atoms or Na atoms. The K-doped g-C3N4 showed

8

enhanced visible light absorption and strong oxidation ability. K atoms intercalated into the

9

g-C3N4 interlayer served as delivery channel via bridging the layers, which is beneficial to the

10

charge carriers separation and transfer between adjacent layers. However, the Na atoms

11

existing in the caves of CN plane increased the in-planar electron density and caused high

12

carriers recombination. As a result, K-intercalated g-C3N4 exhibited high visible

13

photocatalytic performance and Na-doped g-C3N4 showed decreased activity for NO removal

14

in comparison with g-C3N4. The present work has provided new insights into the

15

understanding of photocatalysis mechanism and doping chemistry. Our present finding opens

16

a new avenue for synthesis of efficient intercalated photocatalysts for applications.

17 18

AUTHOR INFORMATIONS

19

Corresponding Author

20

† These two authors contributed equally to this work.

21

*E-mail: [email protected] (F. Dong). Tel.: +86 23 62769785 605, Fax: +86 23 62769785

22

605.

23

Notes

24

The authors declare no competing financial interest.

25

ACKNOWLEDGEMENTS

26

This research is financially supported by the National Natural Science Foundation of China 23 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

(51478070, 51508356, 51108487).

2

REFERENCES

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

(1) Xu, M.; Liang, T.; Shi, M.; Chen, H. Chem. Rev. 2013, 113, 3766–3798. (2) Low, J.; Cao, S.; Yu, J.; Wageh, S. Chem. Commun. 2014, 50, 10768–10777. (3) Butler, S. Z.; Hollen, S. M.; Cao, L.; Cui, Y.; Gupta, J. A.; Gutiérrez, H. R.; Heinz, T. F.; Hong, S. S.; Huang, J.; Ismach, A. F.; Johnston-Halperin, E.; Kuno, M.; Plashnitsa, V. V.; Robinson, R. D.; Ruoff, R. S.; Salahuddin, S.; Shan, J.; Shi, L.; Spencer, M. G.; Terrones, M.; Windl, W.; Goldberger, J. E. ACS Nano 2013, 7, 2898–2926. (4) Tan, C.; Zhang, H. Chem. Soc. Rev. 2015, 44, 2713–2731. (5) Sajjad, M.; Morell, G.; Feng, P. ACS Appl. Mater. Interfaces 2013, 5, 5051−5056. (6) Sun, Z.; Chang, H. ACS Nano 2014, 8, 4133–4156. (7) Zhao, Z.; Sun, Y.; Dong, F. Nanoscale 2015, 7, 15–37. (8) Song, X.; Hub, J.; Zeng, H. J. Mater. Chem. C 2013, 1, 2952–2969. (9) Wang, X. C.; Maeda, K.; Thomas, A.; TakaNae, K.; Xin, G.; Carlsson, J. M.; Domen, K.; Antonietti, M. Nat. Mater. 2009, 8, 76–80. (10) Wang, X.; Blechert, S.; Antonietti, M. ACS Catal. 2012, 2, 1596–1606. (11) Cao, S. W.; Yu, J. G. J. Phys. Chem. Lett. 2014, 5, 2101–2107. (12) Dong, F.; Wang, Z.; Li, Y.; Ho, W.-K.; Lee, S. C. Environ. Sci. Technol. 2014, 48, 10345–10353. (13) Shi, H.; Chen, G.; Zhang, C.; Zou, Z. ACS Catal. 2014, 4, 3637–3643. (14) Xu, L.; Xia, J.; Xu, H.; Yin, S.; Wang, K.; Huang, L.; Wang, L.; Li, H. J. Power Sources 2014, 245, 866–874. (15) Dong, F.; Wang, Z.; Sun, Y.; Ho, W.-K.; Zhang. H. J. Colloid Interface. Sci. 2013, 401, 70–79. (16) Dong, F.; Ou, M.; Jiang, Y.; Guo, S.; Wu, Z. Ind. Eng. Chem. Res. 2014, 53, 2318–2330. (17) Yu, J.; Wang, S.; Low, J.; Xiao, W. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. (18) Yuan, Y.-P.; Cao, S.-W.; Liao, Y.-S.; Yin, L.-S.; Xue, C. Appl. Catal. B. Environ. 2013, 140-141, 164–168. (19) Zhang, G.; Zang, S.; Wang, X. ACS Catal. 2015, 5, 941–947. (20) Bai, X.; Zong, R.; Liu, D.; Zhu, Y. Appl. Catal. B. Environ. 2014, 147, 82–91. (21) Dong, F.; Zhao, Z. W.; Xiong, T. Ni, Z. L.; Zhang, W. D.; Sun, Y. J.; Ho, W. K. ACS Appl. Mater. Interface 2013, 5, 11392–11401. (22) Yang, S.; Gong, Y.; Zhang, J.; Zhan, L.; Ma, L.; Fang, Z.; Vajtai, R. Wang, X.; Ajayan, P. M. Adv Mater. 2013, 25, 2452–2456. (23) Zhang, G.; Zhang, M.; Ye, X.; Qiu, X.; Lin, S.; Wang, X. Adv. Mater. 2014, 26, 805–809. (24) Gao, H.; Yan, S.; Wang, J.; Zou, Z. Dalton Trans. 2014, 43, 8178-8183 (25) Wu, P.; Wang, J.; Zhao, J., Guo, L.; Osterloh, F. E. Chem. Commun. 2014, 50, 15521-15524. (26) Liu, G.; Niu, P.; Sun, C.; Smith, S. C.; Chen, Z.; Lu, G. Q.; Cheng, H.-M. J. Am. Chem. Soc. 2010, 132, 11642–11648. (27) Wang, Y.; Di, Y.; Antonietti, M.; Li, H.; Chen, X.; Wang, X. Chem. Mater. 2010, 22, 5119–5121. 24 ACS Paragon Plus Environment

Page 24 of 26

Page 25 of 26

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 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 39 40 41

(28) Dong, G.; Zhao, K.; Zhang, L. Chem. Commun. 2012, 48, 6178–6180. (29) Yan, S. C.; Li, Z. S.; Zou, Z. G. Langmuir 2010, 26, 3894–3901. (30) Chen, X.; Zhang, J.; Fu, X.; Antonietti, M.; Wang, X. J. Am. Chem. Soc. 2009, 131, 11658–11659. (31) Zhang, G.; Huang, C.; Wang, X. Small 2015, 11, 1215–1221. (32) Ding, Z.; Chen, X.; Antonietti, M.; Wang, X. ChemSusChem 2011, 4, 274–281. (33) Kanetani, K.; Sugawara, K.; Sato, T.; Shimizu, R.; Iwaya K.; Hitosugi, T.; Takahashi, T. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 19610–19613. (34) Wehenkel, D. J.; Bointon1, T. H.; Booth, T.; Bøggild, P.; Craciun1, M. F.; Russo, S. Sci. Rep. 2015, 5, 7609. (35) Oshima, T.; Lu, D.; Ishitani, O.; Maeda, K. Angew. Chem. Int. Ed. 2015, 54, 1–6. (36) Gao, H.; Yan, S.; Wang, J.; Huang, Y. A.; Wang, P.; Li, Z.; Zou, Z. Phys. Chem. Chem. Phys. 2013, 15, 18077–18084. (37) Wu, M.; Yan, J.-M.; Tang, X.; Zhao, M.; Jiang, Q. ChemSusChem 2014, 7, 2654–2658. (38) Zhang, M.; Bai, X.; Liu, D.; Wang, J.; Zhu, Y. Appl. Catal. B. Environ. 2015, 164, 77–81. (39) Kresse, G.; Furthmüller, J. Phys. Rev. B. 1996, 54, 11169–11186. (40) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865−3868. (41) Kresse, G.; Joubert, D. Phys. Rev. B. 1999, 59, 1758–1775. (42) Grimme, S. J. Comput. Chem. 2006, 27, 1787–1799. (43) Ma, X.; Lv, Y.; Xu, J.; Liu, Y.; Zhang, R.; Zhu, Y. J. Phys. Chem. C. 2012, 116, 23485−23493. (44) Gracia, J.; Kroll, P. J. Mater. Chem. 2009, 19, 3013−3019. (45) Speight, J. G. Lange’s Handbook of Chemistry, 16th ed.; McGraw-Hill: New York, 2005, pp 1.155. (46) Becke, A. D.; Edgecombe, K. E. J. Chem. Phys. 1990, 92, 5397–5403. (47) Park, H.; Choi, W. J. Phys. Chem. B. 2003, 107, 3885−3890. (48) Dong, F.; Li, Q. Y.; Sun, Y. J.; Ho, W.-K. ACS Catal. 2014, 4, 4341−4350. (49) Cao, S.; Low, J.; Yu, J.; Jaroniec, M. Adv. Mater. 2015, 27, 2150–2176. (50) Zhang, G.; Zhang, J.; Zhang, M.; Wang, X. J. Mater. Chem. 2012, 22, 8083–8091. (51) Sharma, J.; Gora, T.; Rimstidt, J.; Staley, R. Chem. Phys. Lett. 1972, 15, 232–235. (52) Park, K. H.; Kim, B. H.; Song, S. H.; Kwon, J.; Kong, B. S.; Kang, K.; Jeon, S. Nano Lett. 2012, 12, 2871−2876. (53) Zhao, Y.; Zhao, F.; Wang, X.; Xu, C.; Zhang, Z.; Shi, G.; Qu, L. Angew. Chem. Int. Ed. 2014, 53, 13934–13939. (54) Murphy, A. B. Sol. Energy Mater. Sol. Cells 2007, 91, 1326–1337. (55) Lin, H.; Ding, L.; Pei, Z.; Zhou, Y.; Long, J.; Deng, W.; Wang, X. Appl. Catal. B. Environ. 2014, 160–161, 98–105. (56) Liu, G.; Wang, T.; Zhou, W.; Meng, X.; Zhang, H.; Liu, H.; Kako, T.; Ye, J. J. Mater. Chem. C. 2015, 3, 7538-7542. (57) Dong, F.; Zheng, A.; Sun, Y.; Fu, M.; Jiang, B.; Ho, W.-K.; Lee, S. C.; Wu, Z. CrystEngComm 2012, 14, 3534–3544.

42

(58) Wang, X.; Feng, Z.; Shi, J.; Jia, G.; Shen, S.; Zhou, J.; Li, C. Phys. Chem. Chem. Phys

43 44

2010, 12, 7083–7090. (59) Ishii, T.; Kato, H.; Kudo, A. J. Photoch. Photobio. A. 2004, 163, 181–186. 25 ACS Paragon Plus Environment

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 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3

Page 26 of 26

Graphical Abstract g-C3N4

K-intercalated g-C3N4

K intercalation

Bridging the layers Charge redistribution Tuning band structure Enhanced photocatalysis

4

26 ACS Paragon Plus Environment