You are on page 1of 31

Journal Pre-proof

Oxygen-doped crystalline carbon nitride with greatly extended


visible-light-responsive range for photocatalytic H2 generation

Guoqiang Zhang (Conceptualization) (Data curation) (Investigation)


(Methodology) (Software) (Writing - original draft), Yangsen Xu
(Visualization) (Writing - review and editing), Chuanxin He
(Supervision) (Software), Peixin Zhang (Supervision) (Software),
Hongwei Mi (Conceptualization) (Funding acquisition) (Project
administration) (Supervision) (Writing - review and editing)

PII: S0926-3373(20)31051-1
DOI: https://doi.org/10.1016/j.apcatb.2020.119636
Reference: APCATB 119636

To appear in: Applied Catalysis B: Environmental

Received Date: 14 September 2020


Revised Date: 7 October 2020
Accepted Date: 12 October 2020

Please cite this article as: Zhang G, Xu Y, He C, Zhang P, Mi H, Oxygen-doped crystalline


carbon nitride with greatly extended visible-light-responsive range for photocatalytic H2
generation, Applied Catalysis B: Environmental (2020),
doi: https://doi.org/10.1016/j.apcatb.2020.119636
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Oxygen-doped crystalline carbon nitride with greatly extended visible-light-responsive

range for photocatalytic H2 generation

Guoqiang Zhang,a Yangsen Xu,c Chuanxin He,a,b Peixin Zhang,a,b Hongwei Mi*,a,b
a
College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen,

Guangdong, 518060, PR China


b
Guangdong Flexible Wearable Energy and Tools Engineering Technology Research Centre

Shenzhen University, Shenzhen, Guangdong, 518060, PR China


c
Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen 518060, PR China

of
*Correspondence E-mail: milia807@szu.edu.cn

ro
Graphical Abstract
-p
re
lP
na
ur
Jo

Highlights
 Constructing a series of O-doped K+ implanted CCN with narrowed bandgap (2.71-1.62
eV) for the first time;

 Adopting oxygen doping strategy to activate the n→π* transition and thus enhanced
visible light and even near-infrared light harvesting;

1
 An approximately 45- and 10- times enhancement than pristine CN and CCN for H2
generation at λ > 500 nm;

 Excellent photoactive up to 650 nm owns prominent advantages over most CCN;

ABSTRACT

Crystalline carbon nitride (CCN) materials with photoresponse of more than 600 nm are rare.

Here, we successfully prepared a series of oxygen doped K+ implanted CCN (KCN) with

of
narrowed bandgap (2.71-1.62 eV) for the first time. Compared with most of the O-doped

amorphous CN, the optical absorption can only reach 500 nm, and a few more than 500 nm,

ro
our oxygen doping strategy is adopted to activate more n→π* transitions to enhance visible
-p
light and even near-infrared light harvesting. This active-optimized O-doped KCN accounts for

45- and 10- times promotion than pristine CN and KCN in H2 generation under λ > 500 nm.
re
Most importantly, its maximum active wavelength is up to 650 nm, which has obvious

advantages than most CCN-based photocatalysts in the utilization of solar energy. This
lP

excellent solar capture and H2 production is attributed to the activated n→π* electron transition,

and high crystallinity and conduction band (CB) position caused by oxygen doping.
na

Keywords: crystalline carbon nitride, solar capture, n→π* electron transition, oxygen doping,
ur

photocatalytic H2 generation
Jo

1. Introduction

Polymer carbon nitride (CN) has become a research hotspot for visible-light-driven H2

production due to its advantages of easy synthesis, flexible modification, non-toxicity and

excellent chemical stability.[1] In traditional thermal polymerization, because of the

insufficient mobility of the reaction intermediates, the CN is not fully polymerized and leads to
2
an amorphous melon-based structure.[2] Great progress has been made in extending the

photoresponse of CN. For example, Wang's group reported a new-type triazine-based CN

structure with light absorption edge up to 735 nm[3]. Liu et al. realized the near-infrared

photoresponse by sensitizing protonated two dimensional (2D) CN with Chlorin e6[4]. Guo's

group developed a carbon quantum dot/CN hybrid photocatalyst that responds to infrared light

irradiation[5]. However, these amorphous structures result in weak conjugated aromatic ring

system and the difficulty of charge transfer on the polymer plane. As a result, the CN only

exhibits the sluggish photogenerated carrier separation.[6,7] Recently, crystalline carbon nitride

of
(CCN) has been constructed via a solid-salt or molten-salt assisted approach and emerged as a

promising candidate for driving solar energy conversion.[8] First, the fuller polymerized

ro
polymer heptazines can provide a unblocked charge transfer channel on 2D π-conjugated plane,
-p
and availably accelerate the intralayer charge transfer and exciton dissociation. Besides, the

shorter π-π stacking distance than amorphous melon-based CN, improves lateral charge transfer
re
and exciton dissociation. Therefore, many active top-level CCN have been constructed,

including potassium poly(heptazine imides),[9] K+ implanted CCN,[10-13] defective


lP

CCN,[14,15] and K-Li CCN.[16-18] However, their deficient light response (less than 600 nm)

severely restricts the utilization of sunlight, while the visible light and near-infrared light
na

account for 45% and 50% of the solar energy, respectively. Further extending the light

harvesting of CCN is a worthwhile mission.


ur

2. For energy band structure of CCN, the valence band (VB) is principally composed
Jo

of the 2p orbital of N atoms, while the hybridized C 2p and N 2p orbitals contribute

to conduction band (CB), forming π* interactions.[19] The intrinsic absorption band

derives from the π→π* electron transition with the sp2 hybridization of C and N in

CCN skeleton.[20] This transition intensity increases with the extension of the conjugated

aromatic ring system and the tighter and better packing of the joint heptazine system.

3
Therefore, in sharp contrast with CN, the CCN presents a significant redshift in visible light

absorption. Other n→π* electron transition in the visible region greater than 500 nm,

involving the lone pairs of the edge nitrogen atoms, unfortunately, is spatially forbidden on

perfectly symmetric and planar units and hard to activate.[20-24] Besides, the n→π*

transitions are intrinsically much less photoactive than π→π* transitions due to the huge

difference in orbital electron density.[20-24] It has thus caughted our interest to exploit a

appropriate policies to activate more efficient n→π* transitions with an new absorption

band larger than 500 nm to reinforce its overall ability for solar energy utilization. If an

of
oxygen atom substitutes a nitrogen atom in the heptazine unit, a lone pairs will be

generated. In addition, O incorporation into the heptazine network may activate the

ro
distribution of the nonbonding electrons on the N sites.[24] More importantly, it could
-p
lead to the symmetry changes in forms of C and N atoms on non-coplanar structure, which

fully activate this transition in structural distorted frameworks by a protocol of reducing the
re
lengthways dimensionality.[21,22] Due to such variations, a unique optical-absorption

property is predicted to occur for more n→π* transitions in O-doped heptazine units, and
lP

it could become stronger with the increase of oxygen doping. In consequence, this

oxygen doping strategy is expected to associate with more opportunities for n→π*
na

transitions to enhance visible light and even near-infrared light harvesting.


ur

3. The light absorption of most O-doped amorphous CN is less than 500 nm[25-28]. The O-

doped CN was firstly synthesized by a facile H2O2 hydrothermal approach, and the O-
Jo

doping could induce intrinsic electronic and band structure modulation, resulting in its

absorbance edge up to 498 nm[28]. Huang et al. reported a facile precursor pre-treatment

method to fabricate CN with simultaneous novel porous network, controllable O-doping

and light absorption to 480 nm, by forming hydrogen-bond induced supramolecular

aggregates [27]. Zeng et al. have been facilely prepared one-dimensional O-doped CN

4
nanorods by direct calcination of hydrous melamine nanofibers precipitated from aqueous

solution of melamine[25]. The markedly higher light absorption ranges up to 700 nm for O-

doped CN was reported via one step homogeneous polymerization of urea and oxalic acid

(OA)[29]. These results show that OA has obvious advantages as an oxygen doping source.

Inspired by this result, we chose OA as the oxygen doping source.

4. Herein, heptazine-based KCN has been built by a potassium chloride (KCl) solid-salt-

assisted approach. OA was deliberately chosen as the oxygen doping source to provide

oxygen doping. Thus, a series of O-doped KCN (OKCN) photocatalysts with narrow

of
bandgaps of 2.71-1.62 eV were successfully constructed via altering the quality of OA.

ro
Compared with most of the O-doped amorphous CN, the optical absorption can only reach

500 nm[25-28], and a few more than 500 nm[29], our oxygen doping strategy is adopted to
-p
activate more n→π* transitions to enhance visible light and even near-infrared light
re
harvesting. The optimized activity is approximately 45- and 10- times better than CN and

KCN for H2 production under λ > 500 nm. In addition, this is the first report of O-doped
lP

CCN, and its photoactive wavelength up to 650 nm owns prominent advantages over most

CCN. This excellent photoactive is attributed to the successfully activated new absorption
na

bands of n→π* electron transition in O-doped heptazine units. Our findings provide a new

way to achieve a wide spectral response through doping to activate the n→π* electron
ur

transition.
Jo

5. Experimental Section

2.1 Chemicals and Materials

Urea (AR, 99.0%), triethanolamine (TEOA, AR, 98.0%), KCl (AR, 99.5%), oxalic acid

(OA, 99.0%) and H2PtCl6.6H2O (AR, Pt≥37.5%) were purchased from Aladdin Reagent

Company. All chemicals were reagent grade and used without further purification.

5
2.2 Preparation of CN

Typically, the 20 g of urea powder is put into a 50 ml of crucible with a cover and

then heat to 550°C at a rate of 0.5 °C/min in a muffle furnace and maintain at this

temperature for 3 h. CN is obtained after cooling down to room temperature.

2.3 Preparation of KCN and OKCN

Typically, the 20 g of urea powder mechanical mixing with the 20 g of KCl and 0-

5 g of OA is put into a 50 ml of crucible with a cover and then heat to 600°C at a rate of

0.5 °C/min in a muffle furnace and maintain at this temperature for 6 h. The products

of
were obtained after cooled to room temperature, washed with plenty of water and dried

ro
at 60°C under vacuum. According to the feed quality of OA, we denote the samples as

OKCN x (x = 1-5), where x is the feed quality of OA. KCN is synthesized under the
-p
condition similar to OKCN, except that OA is not added
re
2.4 Characterizations

Transmission electron microscope (TEM) images were taken using a FEI Tecnai
lP

G2 F20 operated at 200 kV. The crystalline structure was recorded by using an X-ray

diffractometer (XRD) (Ultima IV) with Cu K radiation (λ = 1.54056 A). Brunauer-


na

Emmett-Teller (BET) specific surface area was measured using an ASAP2460 Surface

Area and Pore Size Analyzer. The UV-Vis absorption spectra were recorded on a UV-
ur

3600 UV-Vis-NIR scanning spectrophotometer (Shimadzu). Fluorescence emission

spectra were recorded on a LabRAM HR Evolution spectrograph. X-ray photoelectron


Jo

spectrum (XPS) analyses were performed on an ESCALAB 250Xi spectrometer with an

Al-Kα (1486.6 eV) achromatic X-ray source. Transient state fluorescence spectra were

recorded on an Edinburgh instruments FS5 fluorescence spectrometer. Inductively

coupled plasma spectrometry (ICP-OES) was performed on Ultima 2, Horiba. The solid-

state 13C NMR spectra were recorded on a Bruker Avance II instrument in cross-polarization

6
magic-angle spinning sequence mode. The Fourier transform infrared (FTIR) spectra were

recorded on a Nicolet iz10 spectrometer.

2.5 Photocatalytic H2 production Measurements

The 50 mg of samples added with H2PtCl6 (3 wt% Pt) is placed into a 50 mL of

TEOA solution (20 vol%) in a closed gas circulation system. The visible-light

irradiations were obtained from a 300 W Xe lamp (Perfect Light, PLS-SXE300) with a

UVCUT-420 or UVCUT-500 nm filter. All tests are controlled at 15℃ by circulating

condensate. The evolved gases are detected in situ by using an online gas chromatograph

of
(GC9790II, Fuli) equipped with a thermal conductivity detector (TCD).

ro
2.6 AQE calculations of H2 production

The 50 mg of samples added with H2PtCl6 (3 wt% Pt) is placed into a 50 mL of


-p
TEOA solution (20 vol%) in a closed gas circulation system. The catalyst solution was
re
irradiated by a 300W Xe lamp applying bandpass filters (center at 400, 420, 475, 500,

550, 600 and 650 nm) for 2 h. The average intensity of irradiation is determined by an
lP

FZ-A spectroradiometer (Photoelectric Instrument Factory of Beijing Normal

University). The apparent quantum efficiency (AQE) was calculated from equation:
na

2 × 𝑡ℎ𝑒 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑒𝑣𝑜𝑙𝑣𝑒𝑑 𝐻2 𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒𝑠


AQE = × 100%
𝑡ℎ𝑒 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑖𝑛𝑐𝑖𝑑𝑒𝑛𝑡 𝑝ℎ𝑜𝑡𝑜𝑛𝑠

2.7 Surface photovoltage measurements


ur

Surface photovoltage (SPV) measurement system consisted of a source of


Jo

monochromatic light, a lock-in amplifier (SR 830-DSP) with a light chopper (SR 540)

and a sample chamber. Monochromatic light is provided by a 300 W Xe lamp (PLS-

SXE 300, Beijing Trusttech Co. Ltd, China) and a monochromator (SBP500, Zolix). All

measurements are operated at room temperature and under ambient pressure and samples

are not pretreated prior to the SPV measurement.

2.8 IPCE and photocurrent measurement


7
The 10 mg of samples is dispersed in a 100 mL of 0.2 mg/mL I2/acetone solution

under ultrasonic treatment. A two-electrode process is used to deposit the samples at the

applied potential of 30 V for 5 min. FTO glass substrates with the coated area about 1 ×

3 cm2 is used for both electrodes. Then, the deposited electrode is dried at 200°C for 30

min to remove I2 residues. A conventional threeelectrode process is used to investigate

the photoelectrochemical properties of samples in a quartz cell. An FTO photoanode

deposited samples, Hg/HgCl2, and Pt foil electrode acts as the working electrode,

reference electrode, and counter electrode, respectively. A 0.5 M of Na2SO3 aqueous

of
solution is used as the electrolyte. The photoanode is illuminated by a 300 W Xe lamp

with a monochromator or UVCUT-500 nm filter. The illuminated area was 1 × 1 cm2.

ro
The IPCE (Incident Photon-to-Current Efficiency) was calculated as the following

equation: -p
1240 × photocurrent density (mA cm−2 )
re
IPCE % = × 100
6. wavelength nm × photon flux (mW cm−2 )
lP

2.9 The simulation of frontier molecular orbitals

The distribution of the frontier orbitals was carried out using the density functional theory
na

(DFT) package of DMol3 program. All electron core treatment was used for all atoms. The self-

consistent field tolerance was set at 1.0 × 10−5 Ha. According to the optimized molecular
ur

structure, the spatial distributions of HOMO (highest occupied molecular orbital) and LUMO

(lowest unoccupied molecular orbital) could be acquired.


Jo

3 Results and discussion

3.1 Preparation process, absorption properties and electronic structure

Adopting KCl as a solid template to afford an inter-crystal confined space to

effectively induce the KCN directional growth, its crystallinity can be greatly promoted.

A series of OKCN with high crystallinity were successfully prepared by

8
polycondensation of urea, OA and KCl in one pot for the first time. According to the

feed quality of OA, we denote the samples as OKCN x (x = 1-5), where x is the mass of

OA. The synthesis condition of KCN is similar to that of OKCN except that OA is added.

The OA alone or OA and KCl did not produce products due to the complete oxidation

and decomposition of OA at high temperature in the air.

Narrow bandgap engineering is realized by activating the n→π* electron transition from

O-doped heptazine unit. The possible polycondensation process of heptazine-based OKCN is

shown in Figure 1a. It is reported that urea is gradually converted into isocyanic acid, biuret,

of
cyanuric acid, melamine and melem in the process of polycondensation.[30] These processes

favor oxygen doping into melem through condensation reaction of carboxylic acid and amine,

ro
and then the oxygen-containing melem consequently polymerize to OKCN. Figures 1b-d
-p
dispaly the proposed structure of heptazine-based CN, KCN and OKCN. The structure is

composed of one-dimensional amine-linked heptazine-based melon chains, in which K+ ions


re
are intercalated in the interstitial region.[11] The adjacent polymer chains extend into 2D
lP

supramolecular arrays through hydrogen bonds. The colours of these samples can be

adjusted, as shown in Figure 1e, from light yellow to brown, and finally to black, which

vividly reveals that the bandgaps can be regulated in a wide range. The ultraviolet-visible
na

diffuse reflectance spectra (UV-Vis DRS, Figure 1f) indicates that the absorbance range

is greatly expanded as the proportion of OA increases. The OKCN 5 presents the largest
ur

absorption region among all samples, which even capture the light up to near infrared
Jo

region, while pristine CN and KCN only display an optical absorption edge of 460 and

480 nm, respectively. Interestingly, the absorption range of this series of OKCN samples

is much higher than that of the previous O-doped amorphous CN[25-29]. This may be

due to the better crystallinity than CN, resulting in greater conjugated structures. The

intrinsic absorption band originates from the π→π* electron transition of the sp2

hybridization of C and N. It is clearly red-shifted by 0.21 eV, and the bandgap is

9
decreased from 2.71 eV (KCN) to 2.50 eV (OKCN 4) according to Tauc plots (Figure

1g). Moreover, an obvious new absorption band of the n→π* electron transition is

observed in the region from 500 nm to 900 nm, and it becomes stronger with the increase

of oxygen doping. It is worth noting that in sample OKCN 5, the n→π* transition is

strong enough to make the VB level continuous, so its bandgap narrows sharply to 1.62

eV. This new absorption band activated by oxygen doping can utilize visible light with

wavelengths longer than 500 nm. The VB-XPS analyses (Figure S1) indicate that the VB

potential of OKCN 2 and OKCN 3 is 1.51 and 1.45 eV below the Fermi level, which is 0.15 eV

of
and 0.21 eV higher than that of KCN (1.66 eV), respectively, indicating the n→π* electron

transition raises the VB position. Furthermore, according to the changes of bandgap and VB

ro
position, it is estimated that no obvious change of CB position in KCN and OKCN samples.
-p
Therefore, our narrow bandgap strategy, which maintains a high CB position and a strong

reduction ability of photogenerated electrons, is beneficial to photocatalytic H2 production. The


re
yields of the prepared samples are shown in Figure S2. KCN exhibits the highest yield, while
lP

the yields of OKCN decreased with the addition of OA. All the samples present similar thermal

stability (TG, Figure S3). They begin to decompose slowly at around 500℃ and quickly at 650℃.
na
ur
Jo

10
of
ro
-p
re
lP
na
ur

Figure 1. The possible polycondensation process of OKCN sample (a). The proposed molecular
Jo

structure diagram of CN (b), KCN (c) and OKCN (d). The atoms of C, N, K and O are in brown,

light blue, purple and orange, respectively. The optical images (e), UV-Vis DRS spectra (f) and

Tauc plots of the transformed Kubelka-Munk function vs. the energy (g) of CN, KCN and

OKCN x (x = 1-5).

11
3.2 Morphology characterization

The morphologies are characterized by scanning electron microscopy (SEM) and

transmission electron microscopy (TEM). The representative SEM and TEM images are shown

in Figures S4-8 and Figures 2a-b, which presents the different morphologies before and after

K+ ions intercalation and oxygen doping. CN exhibits a typical nanosheets shape, while KCN

and OKCN emerges as a 100-500 nm nanoparticles feature. Even a few microns of chunks are

observed in OKCN 5 sample. The characteristic lattice fringe space about 1.105 nm is indexed

to (1 1 0) facet (Inset of Figure 2b), while no lattice fringe of CN can be observed in the Inset

of
of Figure S7. The bright spots shown in the selected area electron diffraction (SAED) pattern

(Figures 2c and S8c), further explain the good crystallinity of OKCN and KCN. OKCN 2

ro
exhibits a similar SAED pattern to KCN. The spots in the center rings give the lattice distances
-p
of approximately 1.105 (marked in red) and 0.867 nm (marked in orange), belonging to the (1

1 0) and (0 2 0) facets, respectively. Other spots with lattice distance of 0.316 nm (marked in
re
light blue) are assigned to the (0 0 2) facet, corresponding to the interlayer spacing. The element
lP

mapping (Figures 2e-h and S8e-g) demonstrates the uniform distribution of C, N, O and K in

the entire selected area. The Brunauer-Emmett-Teller (BET) surface area are analysed by

using N2 adsorption-desorption isotherm, as shown in Figure S9. The BET surface areas
na

of CN, KCN and OKCN 2 are estimated to be 72.4, 38.6 and 45.4 m2/g, respectively.
ur
Jo

12
of
ro
Figure 2. The SEM (a), TEM images (b) and FFT pattern (c) of OKCN 2. Inset in (b) is the

high-resolution TEM image. Dark-field TEM image (d) and the corresponding elemental

mappings of C, N, O and K distribution (e-h).


-p
re
3.3 Structure characterization

The elemental analysis (Table S1) indicates that the atomic mole ratios of C:N are
lP

0.66 and 0.71 in CN and KCN, respectively, and these values are close to that of melon

(0.67). With the increase of OA, the mole ratios experiences a gradual promotion from 0.73
na

(OKCN 2) to 0.77 (OKCN 3) and 0.81 (OKCN 5). In addition, the carbon element does not

change significantly, while nitrogen content decreases, indicating that oxygen may replace
ur

nitrogen in KCN framework. The content of K element in KCN and OKCN is 8.68-7.12

wt%, as calculated via inductively coupled plasma (ICP, Table S2) analysis. Figure 3a is
Jo

the typical X-ray diffraction (XRD) patterns of CN, KCN and OKCN. The KCN presents three

diffraction peaks at 8.0°, 10.2° and 28.2°, which are attributed to the (1 1 0), (0 2 0) and (0 0 2)

facets, respectively.[11] As the proportion of OA increases, the peaks at 28.2° slightly shift to

a low angle due to a minor increase in interlayer spacing. In sharp contrast to KCN, OKCN

exhibits relatively weakened and broadened peaks upon OA incorporation. In particular, the

13
weak low-angle peaks at 8.0° and 10.2° vanishes in OKCN 5 sample. The full width at half

maxima (FWHM) values of XRD can reflect the crystallinity of the materials. As shown in

Table S3, the crystallinity of KCN and OKCN is significantly improved compared with pristine

CN. The feature structure is analysed through FTIR spectra to inquiry the effect of OA

incorporation, as shown in Figure 3b. The signal at 811 cm−1 belongs to the out-of-plane

bending of the heptazine rings, while the fingerprint peaks at 1200-1700 cm−1 is characteristic

of stretching and bending modes of the conjugated CN heterocycles.[31,32] Furthermore, KCN

and OKCN demonstrates new band at 2166 cm−1, which is assigned to the stretching vibration

of
of cyano units (C≡N), probably deriving from the decomposition or incomplete polymerization

of heptazine groups.[11,33] The characteristic mode of the C-N vibration appears at 993 and

ro
915 cm−1 in KCN and OKCN, and the new peak at 1153 cm−1 can be classified as the N-H
-p
bending vibration. These modes gradually weaken with the successful copolymerization of OA

into KCN framework.


re
lP
na
ur
Jo

Figure 3. The XRD patterns (a) and FTIR spectra (b) of CN, KCN and OKCN.

13
C cross-polarization (CP)-MAS solid-state NMR spectrum is carried out to analyses the

possible chemical structures. As shown in Figure 4a, two distinct resonance peaks at 155.5 and

161.2 ppm could be assigned to the C atoms in the C-N3 and CN2-(NHx) units,[34-36] revealing

the existence of heptazine structures. The signal at 166.4 ppm belongs to the C atom adjacent

14
to the electron-withdrawing group (-C≡N) in KCN and OKCN. The existence of C, N, O, K

and Cl elements in KCN and OKCN is confirmed by XPS survey spectrum (Figure S10), while

only C, N and O in CN. Trace Cl element is estimated to be less than 0.3 wt% according to XPS

results. High-resolution K 2p XPS (Figure S11) is used to analyze the valence state of K. Two

binding energy at 292.8 and 295.6 eV are assigned to K+ ion,[37] and are different from those

for metallic K (294.7 eV). The doublet separation of K 2p photoelectron lines are 2.8 eV, which

further proves the existence form of K+ ion. The C 1s high resolution XPS spectra can be

described as the superposition of four peaks by Gaussian distribution, located around 288.1,

of
287.0, 286.1 and 284.6 eV, respectively (Figure 4b). The peaks at 288.1 and 284.6 eV are

similar to those in CN, and usually assigned to sp 2-bonded carbon atoms in aromatic N-C=N

ro
species and graphitic carbon, respectively.[6,38]. The weak peak at 287.0 eV probably comes
-p
from C≡N group,[39] which is agreement with the FTIR result. The signal of C-O unit at 286.1

eV experiences gradual increase,[29] which indicates the O-doped heptazine unit may have
re
formed in OKCN framework. As shown in Table S4, the ratio of graphite carbon in C 1s XPS
lP

spectra (Figure 4b) ranges from 0.195 to 0.185. Therefore, the improvement of optical

absorption of OKCN is related to the transition caused by oxygen doping, and has little to do

with the change of graphite carbon. However, compared with CN alone, the proportion of
na

graphite carbon increases sharply after the formation of KCN and OKCN. This may be due to

the improvement of crystallinity and conjugation, leading to an increase in graphite carbon. In


ur

order to obtain more evidence of the structural changes caused by O doping, we adopt the
Jo

Electron Paramagnetic Resonance (EPR) test to detect the spin state of unpaired electrons

(Figure 4c). The pristine CN appears one single Lorentzian line with a g factor of 2.002 due to

the unpaired electron from the π-bonded aromatic rings. The EPR intensity of OKCN 2 is

greatly enhanced, probably because substituting the N atom in the heptazine unit with an O

atom creates one unpaired spin. Interestingly, this unpaired electrons produced by the oxygen

doping can generate the n→π* electron transition and thus greatly expand the light absorption.

15
Based on the characterization results, the possible chemical structure of the O-doped heptazine

unit is proposed in Figure 4d. A part of the sp2 N atoms in the heptazine units are substituted

by O atoms during the copolymerization process of urea and OA.

of
ro
-p
re
lP
na

Figure 4. The solid-state CP/MAS 13C NMR spectra (a), the highresolution C 1s XPS spectra

(b) and EPR spectra (c) of CN, KCN and OKCN. The possible diagram of O-doped heptazine
ur

unit (d).
Jo

3.4 Testing and interpretation of photocatalytic activity

This successfully activated new absorption bands of n→π* electron transition helps to

explore their photocatalytic activity under visible light, even at greater than 500 nm. The visible

light is obtained by irradiating with Xe lamp and using 420 and 500 nm cut-off filters (Figures

S12-13). TEOA solution is adopted as sacrificial agent. Figure 5a shows the structure-

dependent photocatalytic activity. CN exhibits a H2 production rate of 81.4 μmol/h for 50 mg


16
of sample. Its performance increases sharply to 392.2 μmol/h for KCN, which is ~3.8 times

higher than CN. And then it gradually drops with the increase of oxygen doping. Figure 5b is

the photocatalytic activity test after 500 nm cutoff. The weak activity (1.0 and 4.1 μmol/h) may

originate from their band tail absorption in CN and KCN samples. A series of OKCN samples

present outstanding performance, among which OKCN 2 possesses the highest rate with 46.1

μmol/h and accounts for an approximately 45- and 10- times promotion than CN and KCN

under λ > 500 nm. The OKCN 5 sample with the narrowest bandgap of 1.62 eV exhibits quite

weak activity, which possibly due to the formation of recombination centers and the reduction

of
of charge separation efficiency caused by excessive oxygen doping. Figure 5c is the apparent

quantum efficiency (AQE) measurement. The AQE of KCN and OKCN 2 can reach 20.5%

ro
and 10.9% at 420 nm, which is 4.0 and 1.7 times better than CN (~4.1%), respectively. At 500
-p
nm, the AQE of OKCN 2 is 2.9%, while trace H2 is detected for CN and KCN. Even at 650

nm, its AQE is still as high as 0.3%. This red-shift of photoresponse from 475 nm (CN) and
re
500 nm (KCN) to 650 nm can be attributed to the activated n→π* electron transition in OKCN
lP

2. After six photocatalytic cycles (Figure 5d), the photocatalytic property has no obvious

attenuation. In addition, there is no obvious change in XRD patterns (Figure S14a) and XPS
na

survey spectra (Figure S14b) before and after photocatalytic reaction. These results fully

demonstrate the excellent photostability of the OKCN 2 photocatalyst. To assess its

performance objectively, we compare it with the first-class CN and CCN -based photocatalysts
ur

reported so far. As shown in Figure 5e, although the maximum active wavelength of OKCN 2
Jo

is not as good as that of CN with 700 nm[3] or near-infrared light[4,5], it has obvious advantages

over most CCN[9,14,39,40-44] and some CN[45-49].

17
of
ro
-p
re
lP

Figure 5. The H2 production rate of CN, KCN and OKCN x (x = 1-5) under λ > 420 (a) and
na

500 nm (b) irradiations. Wavelength-dependent apparent quantum yield of CN, KCN and

OKCN 2 under monochromatic light irradiation (c). The recycling measurements of the H2
ur

production (λ > 500 nm) over OKCN 2 (d). Comparison of active wavelength with the reported
Jo

first-class CCN and CN-based photocatalysts (e). The 50 mg of samples added with H2PtCl6

(3 wt% Pt) is placed into a 50 mL of 20 vol% TEOA solution. The visible-light

irradiations were obtained from a 300 W Xe lamp with a UVCUT-420 or UVCUT-500

nm filters. All tests are controlled at 15℃ by circulating condensate.

The separation and transfer of photogenerated carriers is investigated by steady-state and

transient-state photoluminescence (PL) spectrum. As shown in Figure 6a, in sharp contrast to a


18
band-to-band emission peak at ~470 nm from the π*→π electron transition, the PL peak red-

shifts to 505 nm in OKCN 2 sample, which is consistent with the narrowed bandgap induced

by n→π* transition. In addition, the significant fluorescence quenching indicates the less

carriers generation in photocatalytic process. OKCN 2 exhibits faster fluorescence decay

(Figure 6b) with an average lifetime of 1.08 ns than CN (7.42 ns) and KCN (2.67 ns). In order

to explain whether oxygen doping acts as a donor or acceptor, we simulate the structure of O-

doped CN and calculate its spatial distributions of HOMO (highest occupied molecular orbital)

and LUMO (lowest unoccupied molecular orbital). As shown in Figure S15, the spatial

of
distribution of HOMO and LUMO is not completely separated, but partially overlapped. The

ro
HOMO is mainly distributed in O-doped triazine structure, while LUMO is populated in O-

doped heptazine unit, probably indicating that oxygen doping plays a role as a donor. In addition,
-p
the shallow donor level exists below the CB, while the deep donor level is close to the VB in

forbidden band, and the electron will transition between the donor level and the CB. According
re
to the transition diagram (Figure 7d), the oxygen doping level is more like the deep donor level
lP

above the VB. The deep donor level introduced by oxygen doping will become the

recombination centers of photogenerated carriers. The sharp decay of fluorescence lifetime after
na

oxygen doping may be due to the introduction of the recombination centers. Similarly, the

fluorescence intensity decreased significantly. This also explains that the activity of OKCN 2

is lower than KCN under λ > 420 nm irradiation.


ur
Jo

19
Figure 6. The steady-state (a) and transient-state PL (b) of CN, KCN and OKCN 2.

Surface photovoltage spectroscopy (SPS) is a powerful tool to characterize the

photoelectric response.[50,51] Figure 7a displays surface photovoltage (SPV) response of CN,

KCN and OKCN 2. A positive SPV response band in the range of 350-500 nm is detected,

which is assigned to the π→π* electron transition. Interestingly, the photoelectric response

wavelength is significantly extended to 650 nm in OKCN 2 caused by the activated n→π*

electron transition, which corresponds to the photocatalytic results shown in Figure 5c. The

IPCE measurement (Figure 7b) clearly shows the maximum response wavelength is close to

of
650 nm. The visible-light photocurrent test (Figure 7c) was further adopted to investigate the

ro
photoresponse greater than 500 nm. When the visible light is continuously switched, a series of

photocurrent signals can be detected. Its intensity matches the result of photocatalysis (λ > 500
-p
nm). Figure 7d presents the mechanism of photocatalytic H2 production. The successfully
re
activated new absorption bands of n→π* electron transition can utilize visible light longer than

500 nm. As a sacrificial agent, triethanolamine (TEOA) can rapidly capture photogenerated
lP

holes and be oxidized to TEOA2+.[52] Simultaneously, the photogenerated electrons can be

easily transferred to the Pt nanoparticles and have enough reduction capacity to react with H2O
na

to generate H2 on the surface of the Pt nanoparticles.


ur
Jo

20
of
ro
-p
re
Figure 7. The SPS (a), IPCE (b) and visible-light photocurrent measurement (c) of CN, KCN
lP

and OKCN 2. The schematic diagram of photocatalytic H2 generation under λ > 500 nm

irradiation (d).
na

4 Conclusions

In summary, we have successfully activated a new absorption band of n→π* electron


ur

transition in oxygen-doped KCN, which greatly promotes the photoresponse from 500 to 650

nm. The O-doped heptazine frameworks are confirmed by elemental analysis, XRD, NMR,
Jo

XPS, EPR, etc., and the effect of n→π* electron transition on photoresponse are also fully

investigated by PL, SPS, IPCE and photocurrent tests. Its optimized activity is 45- and 10- times

higher than CN and KCN for H2 evolution under λ > 500 nm. This narrow bandgap strategy

does not significantly reduce the CB position at the expense of the reduction ability of

photogenerated electrons, so it is very beneficial to photocatalytic H2 production. Most of all,

21
its maximum active wavelength has obvious advantages over most CCN. The results presented

herein provide an attractive strategy to construct CCN with wide spectral response, toward

efficient solar capture and H2 production.

Credit Author Statement

Guoqiang Zhang: Conceptualization, Methodology, Data curation, Software,

Investigation, Writing – Original Draft.

Yangsen Xu: Visualization, Writing - Review & Editing.

Chuanxin He: Supervision, Software.

of
Peixin Zhang: Supervision, Software.

ro
Hongwei Mi: Conceptualization, Supervision, Project administration, Writing - Review

& Editing, Funding acquisition.


-p
All authors contributed to manuscript revision, read and approved the submitted
re
version.
lP

Declaration of interests
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
na

Acknowledgment
ur

This work was jointly supported by the Research Grants for Postdoctor in Shenzhen and

Guangdong Basic and Applied Basic Research Foundation (2019A1515111021), and the
Jo

Natural Science Foundation of China (51874199, 22078200).

Supporting Information

More elemental analysis, ICP, XPS, TG, SEM, TEM and BET data are shown in the

supplementary information.

22
References

[1] X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J.M. Carlsson, K. Domen, M.

Antonietti, A metal-free polymeric photocatalyst for hydrogen production from water under

visible light, Nat. Mater., 8 (2009) 76-80.

[2] F.K. Kessler, Y. Zheng, D. Schwarz, C. Merschjann, W. Schnick, X. Wang, M.J. Bojdys,

Functional carbon nitride materials-design strategies for electrochemical devices, Nat. Rev.

Mater., 2 (2017) 17030.

[3] P. Yang, R. Wang, M. Zhou, X. Wang, Photochemical Construction of Carbonitride

of
Structures for Red-Light Redox Catalysis, Angew. Chem. Int. Ed., 57 (2018) 8674-8677.

[4] Y. Liu, M. He, R. Guo, Z. Fang, S. Kang, Z. Ma, M. Dong, W. Wang, L. Cui, Ultrastable

ro
metal-free near-infrared-driven photocatalysts for H2 production based on protonated 2D g-
-p
C3N4 sensitized with Chlorin e6, Appl. Catal. B-Environ., 260 (2020) 118137.

[5] Y. Guo, P. Yao, D. Zhu, C. Gu, A novel method for the development of a carbon quantum
re
dot/carbon nitride hybrid photocatalyst that responds to infrared light irradiation, J. Mater.
lP

Chem. A, 3 (2015) 13189-13192.

[6] S. Cao, J. Low, J. Yu, M. Jaroniec, Polymeric photocatalysts based on graphitic carbon

nitride, Adv. Mater., 27 (2015) 2150-2176.


na

[7] J. Wen, J. Xie, X. Chen, X. Li, A review on g-C3N4 -based photocatalysts, Appl. Surf. Sci.,

391 (2016) 72-123.


ur

[8] L. Lin, Z. Yu, X. Wang, Crystalline carbon nitride semiconductors for photocatalytic water
Jo

splitting, Angew. Chem. Int. Ed., 58 (2019) 6164-6175.

[9] A. Savateev, S. Pronkin, J.D. Epping, M.G. Willinger, C. Wolff, D. Neher, M. Antonietti,

D. Dontsova, Potassium Poly(heptazine imides) from Aminotetrazoles: Shifting Band Gaps of

Carbon Nitride-like Materials for More Efficient Solar Hydrogen and Oxygen Evolution,

ChemCatChem, 9 (2017) 167-174.

23
[10] L. Tian, J. Li, F. Liang, J. Wang, S. Li, H. Zhang, S. Zhang, Molten salt synthesis of

tetragonal carbon nitride hollow tubes and their application for removal of pollutants from

wastewater, Appl. Catal. B-Environ., 225 (2018) 307-313.

[11] Y. Xu, X. He, H. Zhong, D.J. Singh, L. Zhang, R. Wang, Solid salt confinement effect: An

effective strategy to fabricate high crystalline polymer carbon nitride for enhanced

photocatalytic hydrogen evolution, Appl. Catal. B-Environ., 246 (2019) 349-355.

[12] M. Wu, J. Yan, X. Tang, M. Zhao, Q. Jiang, Synthesis of Potassium-Modified Graphitic

Carbon Nitride with High Photocatalytic Activity for Hydrogen Evolution, ChemSusChem, 7

of
(2014) 2654-2658.

[13] Y. Xu, C. Qiu, X. Fan, Y. Xiao, G. Zhang, K. Yu, H. Ju, X. Ling, Y. Zhu, C. Su, K +-

ro
induced crystallization of polymeric carbon nitride to boost its photocatalytic activity for H 2
-p
evolution and hydrogenation of alkenes, Appl. Catal. B-Environ., 268 (2020) 118457.

[14] W. Ren, J. Cheng, H. Ou, C. Huang. M.M. Titirici, X. Wang, Enhancing Visible light
re
Hydrogen Evolution Performance of Crystalline Carbon Nitride by Defect Engineering,
lP

ChemSusChem, 12 (2019) 3257-3262.

[15] V.W.-h. Lau, I. Moudrakovski, T. Botari, S. Weinberger, M.B. Mesch, V. Duppel, J.

Senker, V. Blum, B.V. Lotsch, Rational design of carbon nitride photocatalysts by identification
na

of cyanamide defects as catalytically relevant sites, Nat. Commun., 7 (2016) 12165.

[16] G. Zhang, G. Li, Z.A. Lan, L. Lin, A. Savateev, T. Heil, S. Zafeiratos, X. Wang, M.
ur

Antonietti, Optimizing Optical Absorption, Exciton Dissociation, and Charge Transfer of a


Jo

Polymeric Carbon Nitride with Ultrahigh Solar Hydrogen Production Activity, Angew. Chem.

Int. Ed., 56 (2017) 13445-13449.

[17] Z. Zeng, X. Quan, H. Yu, S. Chen, Y. Zhang, H. Zhao, S. Zhang, Carbon nitride with

electron storage property: Enhanced exciton dissociation for high-efficient photocatalysis,

Appl. Catal. B-Environ., 236 (2018) 99-106.

24
[18] Y. Li, F. Gong, Q. Zhou, X. Feng, J. Fan, Q. Xiang, Crystalline isotype heptazine-/triazine-

based carbon nitride heterojunctions for an improved hydrogen evolution, Appl. Catal. B-

Environ., 236 (2018) 99-106.

[19] A. Du, S. Sanvito, Z. Li, D. Wang, Y. Jiao, T. Liao, Q. Sun, Y.H. Ng, Z. Zhu, R. Amal,

Hybrid graphene and graphitic carbon nitride nanocomposite: gap opening, electron-hole

puddle, interfacial charge transfer, and enhanced visible light response, J. Am. Chem. Soc., 134

(2012) 4393-4397.

[20] G. Zhang, G. Li, Z.A. Lan, L. Lin, A. Savateev, T. Heil, S. Zafeiratos, X. Wang, M.

of
Antonietti, Optimizing optical absorption, exciton dissociation, and charge transfer of a

polymeric carbon nitride with ultrahigh solar hydrogen production activity, Angew. Chem. Int.

ro
Ed., 56 (2017) 13445-13449.
-p
[21] Y. Chen, B. Wang, S. Lin, Y. Zhang, X. Wang, Activation of n→π* transitions in two-

dimensional conjugated polymers for visible light photocatalysis, J. Phys. Chem. C, 118 (2014)
re
29981-29989.

[22] A.B. Jorge, D.J. Martin, M.T. Dhanoa, A.S. Rahman, N. Makwana, J. Tang, A. Sella, F.
lP

Corà, S. Firth, J.A. Darr, H2 and O2 evolution from water half-splitting reactions by graphitic

carbon nitride materials, J. Phys. Chem. C, 117 (2013) 7178-7185.


na

[23] G. Zhang, A. Savateev, Y. Zhao, L. Li, M. Antonietti, Advancing the n→π* electron

transition of carbon nitride nanotubes for H2 photosynthesis, J. Mater. Chem. A, 5 (2017)


ur

12723-12728.
Jo

[24] Y. Li, J. Zhang, Q. Wang, Y. Jin, D. Huang, Q. Cui, G. Zou, Nitrogen-rich carbon nitride

hollow vessels: synthesis, characterization, and their properties, J. Phys. Chem. B, 114 (2010)

9429-9434.

[25] Y. Zeng, X. Liu, C. Liu, L. Wang, Y. Xia, S. Zhang, S. Luo, Y. Pei, Scalable one-step

production of porous oxygen-doped g-C3N4 nanorods with effective electron separation for

excellent visible-light photocatalytic activity, Appl. Catal. B-Environ., 224 (2018) 1-9.

25
[26] W. Fang, J. Liu, L. Yu, Z. Jiang, W. Shangguan, Novel (Na, O) co-doped g-C3N4 with

simultaneously enhanced absorption and narrowed bandgap for highly efficient hydrogen

evolution, Appl. Catal. B-Environ., 209 (2017) 631-636.

[27] Z.-F. Huang, J. Song, L. Pan, Z. Wang, X. Zhang, J.-J. Zou, W. Mi, X. Zhang, L. Wang,

Carbon nitride with simultaneous porous network and O-doping for efficient solar-energy-

driven hydrogen evolution, Nano Energy, 12 (2015) 646-656.

[28] J. Li, B. Shen, Z. Hong, B. Lin, B. Gao, Y. Chen, A facile approach to synthesize novel

oxygen-doped g-C3N4 with superior visible-light photoreactivity, Chem. Commun., 48 (2012)

of
12017-12019.

[29] P. Qiu, C. Xu, H. Chen, F. Jiang, X. Wang, R. Lu, X. Zhang, One step synthesis of oxygen

ro
doped porous graphitic carbon nitride with remarkable improvement of photo-oxidation
-p
activity: Role of oxygen on visible light photocatalytic activity, Appl. Catal. B-Environ., 206

(2017) 319-327.
re
[30] Y. Zhang, J. Liu, G. Wu, W. Chen, Porous graphitic carbon nitride synthesized via direct
lP

polymerization of urea for efficient sunlight-driven photocatalytic hydrogen production,

Nanoscale, 4 (2012) 5300-5303.

[31] J. Zhang, G. Zhang, X. Chen, S. Lin, L. Möhlmann, G. Dołęga, G. Lipner, M. Antonietti,


na

S. Blechert, X. Wang, Co-monomer control of carbon nitride semiconductors to optimize

hydrogen evolution with visible light, Angew. Chem. Int. Ed., 51 (2012) 3183-3187.
ur

[32] M. Shalom, S. Inal, C. Fettkenhauer, D. Neher, M. Antonietti, Improving carbon nitride


Jo

photocatalysis by supramolecular preorganization of monomers, J. Am. Chem. Soc., 135 (2013)

7118-7121.

[33] K. Schwinghammer, B. Tuffy, M.B. Mesch, E. Wirnhier, C. Martineau, F. Taulelle, W.

Schnick, J. Senker, B.V. Lotsch, Triazine-based carbon nitrides for visible-light-driven

hydrogen evolution, Angew. Chem. Int. Ed., 52 (2013) 2435-2439.

26
[34] B.V. Lotsch, M. Döblinger, J. Sehnert, L. Seyfarth, J. Senker, O. Oeckler, W. Schnick,

Unmasking Melon by a Complementary Approach Employing Electron Diffraction, Solid-State

NMR Spectroscopy, and Theoretical Calculations-Structural Characterization of a Carbon

Nitride Polymer, Chem. Eur. J., 13 (2007) 4969-4980.

[35] L. Lin, H. Ou, Y. Zhang, X. Wang, Tri-s-triazine-based crystalline graphitic carbon nitrides

for highly efficient hydrogen evolution photocatalysis, ACS Catal., 6 (2016) 3921-3931.

[36] J.R. Holst, E.G. Gillan, From triazines to heptazines: deciphering the local structure of

amorphous nitrogen-rich carbon nitride materials, J. Am. Chem. Soc., 130 (2008) 7373-7379.

of
[37] A. Ihs, K. Uvdal, B. Liedberg, Infrared and photoelectron spectroscopic studies of ethyl

and octyl xanthate ions adsorbed on metallic and sulfidized gold surfaces, Langmuir, 9 (1993)

ro
733-739.
-p
[38] G. Zhang, J. Zhang, M. Zhang, X. Wang, Polycondensation of thiourea into carbon nitride

semiconductors as visible light photocatalysts, J. Mater. Chem., 22 (2012) 8083-8091.


re
[39] L. Lin, W. Ren, C. Wang, A. Asiri, J. Zhang, X. Wang, Crystalline carbon nitride
lP

semiconductors prepared at different temperatures for photocatalytic hydrogen production,

Appl. Catal. B-Environ., 231 (2018) 234-241.

[40] G. Zhang, M. Liu, T. Heil, S. Zafeiratos, A. Savateev, M. Antonietti, X. Wang, Electron


na

Deficient Monomers that Optimize Nucleation and Enhance the Photocatalytic Redox Activity

of Carbon Nitrides, Angew. Chem., 58 (2019) 14950-14954.


ur

[41] G. Zhang, L. Lin, G. Li, Y. Zhang, A. Savateev, X. Wang, M. Antonietti, Ionothermal


Jo

Synthesis of Triazine-Heptazine Based Co-frameworks with Apparent Quantum Yields of 60%

at 420 nm for Solar Hydrogen Production from "Sea Water", Angew. Chem. Int. Ed., 57 (2018)

9372-9376.

[42] M.K. Bhunia, Y. Kazuo, T. Kazuhiro, Harvesting solar light with crystalline carbon

nitrides for efficient photocatalytic hydrogen evolution, Angew. Chem. Int. Ed., 53 (2015)

11001-11005.

27
[43] Z. Zeng, H. Yu, X. Quan, S. Chen, S. Zhang, Structuring phase junction between tri-s-

triazine and triazine crystalline C3N4 for efficient photocatalytic hydrogen evolution, Appl.

Catal. B-Environ., 227 (2018) 153-160.

[44] H. Ou, L. Lin, Y. Zheng, P. Yang, Y. Fang, X. Wang, Tri-s-triazine-Based Crystalline

Carbon Nitride Nanosheets for an Improved Hydrogen Evolution, Adv. Mater., 29 (2017)

1700008.

[45] Y. Kang, Y. Yang, L.C. Yin, X. Kang, G. Liu, H.M. Cheng, An amorphous carbon nitride

photocatalyst with greatly extended visible-light-responsive range for photocatalytic hydrogen

of
generation, Adv. Mater., 27 (2015) 4572-4577.

[46] Y. Yu, Y. Wei, X. Wang, L. Pei, K. Ding, Surface Engineering for Extremely Enhanced

ro
Charge Separation and Photocatalytic Hydrogen Evolution on g-C3N4, Adv. Mater., 30 (2018)

1705060. -p
[47] Y. Yu, W. Yan, W. Gao, P. Li, X. Wang, S. Wu, W. Song, K. Ding, Aromatic ring
re
substituted g-C3N4 for enhanced photocatalytic hydrogen evolution, J. Mater. Chem. A, 2017,
lP

5, 17199-17203.

[48] J. Li, D. Wu, J. Iocozzia, H. Du, X. Liu, Y. Yuan, W. Zhou, Z. Li, Z. Xue, Z. Lin, Achieving

efficient incorporation of π‐ electrons into graphitic carbon nitride for markedly improved
na

hydrogen generation, Angew. Chem. Int. Ed., 131 (2019) 2007-2011.

[49] J. Liu, Y. Yu, R. Qi, C. Cao, X. Liu, Y. Zheng, W. Song, Enhanced electron separation on
ur

in-plane benzene-ring doped g-C3N4 nanosheets for visible light photocatalytic hydrogen
Jo

evolution, Appl. Catal. B-Environ., 244 (2019) 459-464.

[50] G. Zhang, W. Jiang, S. Hua, H. Zhao, L. Zhang, Z. Sun, Constructing bulk defective

perovskite SrTiO3 nanocubes for high performance photocatalysts, Nanoscale, 8 (2016) 16963-

16968.

[51] L. Kronik, Y. Shapira, Surface photovoltage phenomena: theory, experiment, and

applications, Surf. Sci. Rep., 37 (1999) 1-206.

28
[52] Z. Chai, Q. Li, D. Xu, Photocatalytic reduction of CO2 to CO utilizing a stable and efficient

hetero-homogeneous hybrid system, RSC Adv., 4 (2014) 44991-44995.

of
ro
-p
re
lP
na
ur
Jo

29

You might also like