Next Article in Journal
Mg12O12 and Be12O12 Nanocages as Sorbents and Sensors for H2S and SO2 Gases: A Theoretical Approach
Next Article in Special Issue
Synergy of Au–Pt for Enhancing Ethylene Photodegradation Performance of Flower-like TiO2
Previous Article in Journal
Bimodal Interventional Instrument Markers for Magnetic Particle Imaging and Magnetic Resonance Imaging—A Proof-of-Concept Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Dimensional P-Doped Graphitic Carbon Nitride Tube: Facile Synthesis, Effect of Doping Concentration, and Enhanced Mechanism for Photocatalytic Hydrogen Evolution

1
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
2
School of Materials Science and Engineering, Henan Province International Joint Laboratory of Materials for Solar Energy Conversion and Lithium Sodium Based Battery, Luoyang Institute of Science and Technology, Luoyang 471023, China
3
State Key Laboratory of Advanced Technology for Float Glass, CNBM Research Institute for Advanced Glass Materials Group Co., Ltd., Bengbu 233000, China
4
Silica-Based Materials Laboratory of Anhui Province, Bengbu 233000, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(10), 1759; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12101759
Submission received: 24 April 2022 / Revised: 18 May 2022 / Accepted: 19 May 2022 / Published: 21 May 2022
(This article belongs to the Special Issue Nanocatalyst for Water Splitting)

Abstract

:
P-doped graphitic carbon nitride tubes (P-CNTS) with different P concentrations were successfully fabricated via a pre-hydrothermal in combination with a calcination process under a nitrogen atmosphere. The as-prepared samples exhibited excellent photocatalytic performance with a hydrogen production rate (HPR) of 2749.3 μmol g−1 h−1, which was 17.5 and 6.6 times higher than that of the bulk graphitic carbon nitride (CNB) and graphitic carbon nitride tube (CNT). The structural and textural properties of the P-CNT samples were well-investigated via a series of characterization methods. Compared with the bulk g-C3N4, the tubular structure of the doped samples was provided with a larger specific surface area (SSA) and a relatively rough interior. Besides the above, surface defects were formed due to the doping, which could act as more active sites for the hydrogen production reaction. In addition, the introduction of the P element could effectively adjust the band-gap, strengthen the harvest of visible-light, and boost the effective separation of photogenerated charges. More interestingly, these findings can open up a novel prospect for the enhancement of the photocatalytic performance of the modified g-C3N4.

1. Introduction

Nowadays, we are facing a global energy crisis caused by a growing population, increased demand, and excessive energy consumption. Therefore, it is urgent to explore efficient and sustainable technologies to generate clean and renewable energy. Recently, semiconductor photocatalytic hydrogen production has received extensive attention, which can effectively alleviate the serious energy crisis [1,2,3].
Recently, different semiconductor materials (e.g., TiO2 [4], ZnO [5], CdS [6], Bi2WO6 [7], CaTiO3 [8], ZnIn2S4 [9], g-C3N4 [10,11,12,13,14] etc.) have been reported for photocatalytic hydrogen production. Among these semiconductors, g-C3N4 (CNB) has been widely studied because of its low cost, earth-abundant, non-toxic, stable physicochemical properties, and suitable energy band structure [10,11,15]. However, the CNB sample has the disadvantages of low visible-light utilization and high-photogenerated carrier (electron–hole pairs) quenching rate, which have greatly limited its application in photocatalytic hydrogen production [10,11,12,13,14]. Many strategies have been developed to address these problems such as morphology modulation [16,17], the integration with carbon nanodots [18], and the construction of a heterogeneous structure with other semiconductors (NiO [19], BiOI [20], BiOCl [21], Cu2O [22], MoS2 [23] etc.). However, this modification has still failed to significantly improve the hydrogen production activity of CNB. Compared with the modification of constructing a hetero-structure and element doping (B [24], I [25], O [1,14,26], C [27], Fe [28], P [11,29,30,31,32] etc.) are considered as the most effective strategies to adjust the band-gap and decrease the quenching rate of the photogenerated electron–hole pairs, which led to the improvement in the photocatalytic performance. For example, Zhou et al. successfully synthesized the P-doped g-C3N4 by thermally induced copolymerization [11]. When the design amount of the P element was 2.5 wt% of the precursor, the as-prepared sample had the best hydrogen production performance, where the HPR could reach 506 μmol g−1 h−1, which was 2.9 times higher than that of the bulk g-C3N4. The improved performance can be attributed to the introduction of the P element, which decreased the band-gap, enhanced the visible-light response range, and hindered the recombination of the photogenerated charges. Ran et al. prepared P-doped graphitic carbon nitride nanosheets (PCN-S) for the first time by combining the P-doped and thermal exfoliation methods [30]. The prepared sample had the highest apparent quantum efficiency of 3.56% at 420 nm when the P element was fed at 0.5 wt% of the precursor, which was the highest of the g-C3N4 nanosheet materials. Therefore, the PCN-S sample possessed the highest photocatalytic HPR (1596 μmol g−1 h−1), which was 14.8 times higher than the bulk g-C3N4 (108 μmol g−1 h−1). Guo et al. obtained P-doped tubular g-C3N4 (P-TCN) and the photocatalytic HPR was 570 μmol g−1 h−1, which was 9.5 times higher than the pristine g-C3N4 [32]. The content of the P element in the sample with the best hydrogen production performance was explored to be 0.87 wt% when further analyzed by XPS. The experimental results showed that the introduction of the P element could effectively improve the absorption of visible-light, boost the separation and migration of the photoexcited carriers, and enhance the performance of g-C3N4.
Although many works on the P-doped and morphology modulation have been reported, they only fixed the P concentration and failed to delve into the connection between the doping concentration and the photocatalytic performance [32]. In addition, Putri et al. fabricated O-doped g-C3N4 (O-gC3N4x) with different O-doped concentrations by the hydrothermal method [33], where x = 10, 30, 50, 70, 90, and the O concentration was controlled via the addition of hydrogen peroxide. The experimental results showed that O-gC3N430 possessed the optimal catalytic performance and the hydrogen production performance of the remaining O-doped g-C3N4 was inferior to that of O-gC3N430. This demonstration confirmed the importance of the doping level of the impurity elements, which can greatly affect the hydrogen production performance of the photocatalyst. Based on the above, we intended to vary the P source dosage and investigate the effect of different doping concentrations on the photocatalytic performance of the samples and explore the mechanism of the enhanced photocatalytic performance in depth. In our work, the P-CNT samples with different P contents were synthesized, and the photocatalytic property was evaluated by decomposing aquatic hydrogen. As expected, the as-obtained samples had the optimal photocatalytic performance when the content of P element was 0.9 wt%, with a photocatalytic HPR of 2749.3 μmol g−1 h−1, which was 17.5 and 6.6 times higher than the CNB and CNT samples, respectively. In addition, we also propose a reasonable mechanism to explain the enhanced performance.

2. Materials and Methods

2.1. Materials

Melamine (C3H6N6), triethanolamine (TEOA), and chloroplatinic acid hexahydrate (H2PtCl6·6H2O) were all obtained from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China). Disodium hydrogen phosphate dodecahydrate (Na2HPO4·12H2O) was obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

2.2. Synthesis of Photocatalyst

CNT and P-CNT with different P contents were synthesized via a pre-hydrothermal in combination with the calcination process under a nitrogen atmosphere. The specific preparation flow chart of P-CNT is concisely depicted in Scheme 1. In brief, 1.26 g of melamine was dissolved in 100 mL of deionized water under a water bath at 85 °C, and then different masses of disodium hydrogen phosphate dodecahydrate (Na2HPO4·12H2O) were added to the solution and vigorously stirred for half an hour. Subsequently, the solution was then transferred to the polytetrafluoroethylene liner, then the reactor was screwed down and held at 180 °C for 10 h. After the autoclave was cooled naturally, the mixed solution was centrifuged, washed with deionized water and anhydrous ethanol, and dried in an oven at 60 °C. Finally, the as-obtained solid powder was heated to 500 °C with a heating rate of 2.5 °C/min and held for 4 h under a nitrogen atmosphere. Additionally, the concentrations of P element were designed to be 0.3 wt%, 0.6 wt%, 0.9 wt%, 1.2 wt% and 1.5 wt%, and the corresponding samples were marked as 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT, respectively. For comparison, the CNT photocatalyst was obtained in the absence of Na2HPO4·12H2O via a similar procedure. The CNB sample was acquired through the common air pyrolysis method.

2.3. Characterization

The XRD patterns were gained via a Bruker D8. The SEM images were acquired using a Hitachi S-4800 instrument operating at 5 kV. The TEM photographs were gained from JEOL JEM 1400F (JEOL Ltd., Tokyo, Japan) with an accelerating voltage of 120 kV, and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) photographs were acquired from FEI Talos F200S (Thermo Fisher Scientific, New York, NY, USA). The SSA of samples was determined using a Tristar II 3020 surface area and porosity analyzer (Micromeritics, Norcross, GA, USA) according to the Brunauer–Emmett–Teller (BET) method. The UV–Vis diffuse reflectance spectrum was obtained by a PuXi TU-1901 UV–Vis spectrophotometer (Puxi, Beijing, China), with BaSO4 as the reference. XPS patterns were tested using a Thermo Fisher Scientific ESCALAB 250Xi (Thermo Fisher Scientific, New York, NY, USA) with a monochromatic Al-Kα line source, where the standard C1s peak centered at 284.8 eV was used as the reference. The electron paramagnetic resonance (EPR) patterns were obtained via a Bruker EMX-nano (Bruker, Karlsruhe, Germany).

2.4. Photocatalytic Hydrogen Production Test

The photocatalytic hydrogen production experiment was conducted in a Pyrex top-irradiated reactor connected to a sealed glass gas-circulation system. The light source was a 300 W Xenon lamp (PLS-SEX300C, Perfectlight Technology Co., Ltd., Beijing, China) with a 420 nm cut-off filter. A total of 50 mg of photocatalyst was homogeneously dispersed in a mixture of 80 mL of water and 20 mL of triethanolamine by sonication and stirring. In addition, 500 μL of 1 mg/mL H2PtCl6 solution (1 wt%) was in situ photo-deposited onto the sample surface as a cocatalyst. The reactant system was evacuated to eliminate air from the vessel prior to the light testing. During the photocatalytic reaction, the temperature of the circulating condensate was retained at 5 °C. The generated gases were analyzed online via gas chromatography (GC7900, Agilent, Santa Clara, CA, USA) equipped with a 5Å molecular sieve filled column and a thermal conductive detector (TCD), using high purity nitrogen as the carrier gas.

2.5. Photoelectrochemical Measurement

The photoelectrochemical properties of all samples were tested via the electrochemical workstation connected to a three-electrode system equipped with a counter electrode (Pt plate), a reference electrode (Ag/AgCl), and a working electrode (an FTO glass plate loaded photocatalyst). A neutral Na2SO4 solution (0.5 M, pH = 6.8) was selected as the electrolyte to conduct the charge.
Furthermore, the working electrode was manufactured as follows: First, a mixture consisting of 3 mg photocatalyst, 250 μL deionized water, 250 μL anhydrous ethanol, and 10 μL DuPont D1020 Nafion solutions (5 wt%) was uniformly dispersed with ultrasound for 20 min. Then, 25 μL of the as-prepared mixture was uniformly dropped on an FTO glass with the size 1 × 1 cm2. Finally, the FTO glass was dried under the irradiation of an infrared lamp.

3. Results and Discussion

The X-ray diffraction patterns were investigated to explore the influence of P-doped on the crystal structure of the CNB. As marked in Figure 1, the CNB sample possessed two prominent and sharp diffraction peaks at around 13.2° and 27.7°, which derived from the in-plane structural stacking (100) and cyclical stacking (002) of layers following the c-axis direction [10,34,35], respectively. However, it was noted that the diffraction peak at 13.2° for all the P-CNT samples disappeared and the diffraction peak at 27.7° also became wider and weaker than that of the CNB, mainly due to the hollow structure in the plane and the introduction of the P element, indicating that the P element possesses the effect of reducing the interlayer periodicity due to the mismatch of crystal parameters.
The morphology and microstructure of the CNB and 0.9P-CNT samples were observed via SEM and TEM images. As displayed in Figure 2a, the CNB sample exhibited a typical blocky structure with an irregular and heterogeneous arrangement [36]. After hydrothermal treatment, a supramolecular precursor with a rod-like structure could be obtained (Figure 2b,c), originating from the in situ self-assembly process of melamine and its hydrolyzed cyanuric acid [32]. The rod-shaped precursors were of different sizes and randomly intercrossed. The morphology of the 0.9P-CNT sample is shown in Figure 2d–f, where we can see that it possessed a tubular structure with a relatively rough surface and hollow shape with a diameter of about 1–5 μm. When the rod precursor was calcined, the decomposition may have started from the inside first, and the groups containing the C and N elements in the sample were decomposed and released as gases. Then, the carbon nitride with a tubular structure could be formed under the impact of the flowing N2 atmosphere. The sizes of the as-obtained tubular samples were inhomogeneous due to the inhomogeneous sizes of the precursors.
In addition, Figure 3a shows the TEM image of the 0.9P-CNT sample, which also demonstrates that 0.9P-CNT has a tubular structure with a diameter around 1–3 μm. In addition, the interior of the tubular structure is rough, which can provide more active sites [37,38,39] and promote more efficient charge transfer, further enhancing the photocatalytic performance. More importantly, as seen in Figure 3b, there were a large number of pores in the walls of the tubular structure with a diameter of 30–50 nm, resulting in a special porous structure, which can accelerate the release rate of hydrogen during the hydrogen production process [17,40,41]. Figure 3c,d shows the high-resolution transmission electron microscopy (HRTEM) photographs of the 0.9P-CNT sample, and the inset shows the corresponding selected area electron diffraction images. From the few numbers of lattice stripes in the figure, we can see that the crystallinity of the as-prepared sample was very poor, corresponding to the XRD analysis, while Figure 3d confirms its polycrystalline nature. Figure 3e shows the HAADF-STEM photo of the 0.9P-CNT sample, and Figure 3f–h are the corresponding element mapping images. As illustrated in Figure 3f,g, the C and N elements were uniformly distributed in the tubular sample. Furthermore, Figure 3h confirms that the P element was indeed present in the 0.9P-CNT sample, and the EDX analysis results in Figure 3i provide strong evidence for the doping of the P element into the CNT. More interestingly, inverse Fourier transform images of different orientations could clearly be seen in Figure 4, suggesting crystal distortions in the 0.9P-CNT sample. Specifically, the red symbol T in the figure indicates the defects present in the sample [42], and many symbols of T revealed that the as-prepared sample possessed a large number of surface defects.
The N2 adsorption–desorption curves of the CNB and P-CNT samples are displayed in Figure 5, where all samples exhibited the type IV isotherm, indicating that the as-prepared samples possessed well-defined mesopores from the tubular structure and porous walls [32], which agreed well with the analytical results of the SEM and TEM. Table 1 shows the SSA values of the CNB and P-CNT samples, from which it can be seen that 0.9P-CNT had the highest SSA value. As a consequence, the increased SSA and the amorphous characteristic indicated that more active sites can be generated for the hydrogen production reaction [11,30], resulting in the improvement in the hydrogen production performance.
X-ray photoelectric spectroscopy (XPS) was employed to further investigate the surface elemental composition and chemical status of the P bonded with the other elements. Figure 6a demonstrates the XPS survey spectra, in which peaks belonging to the C, N, and O elements could be observed in the XPS survey spectrum of the 0.9P-CNT sample. The signal of the O element originated from the oxygen adsorbed in the air, however, the absence of the P element detected in the survey spectrum may be due to its low content. To further investigate the interaction of the P element with CNT, the high-resolution XPS tests were performed and the results are displayed in Figure 6b–d. In the C 1s spectrum (Figure 6b) of the 0.9P-CNT sample, two significantly dominant peaks could be found at 284.8 and 288.2 eV, corresponding to the adventitious carbon and the sp2-hybridized carbon containing defects (-N=C-N), respectively [43,44,45]. The N 1s spectrum (Figure 6c) of the 0.9P-CNT sample can be subdivided into four smaller peaks with binding energies of 398.1, 399.4, 400.8, and 404.2 eV [46,47]. The major N 1s peak at 398.1 eV originated from the sp2-hybridized nitrogen atom in the triazine rings (C=N-C), while the peak at 399.4 eV was associated with the N-(C)3 group or the amino group ((C)2-N-H) in the sp3-hybridization. A weak peak at 400.8 eV corresponded to the N-(C)3 groups or a residual amino functional group in the aromatic cycle. In addition, there was an ignorable peak at 404.2 eV from the π excitation [47,48]. The feeble peak at 132.8 eV in Figure 6d was derived from the P–N bond [31,32], further indicating the presence of the P element in the 0.9P-CNT sample. On this basis, the XPS analysis results also strongly demonstrate that P successfully substituted C and formed the P–N bond in the 0.9P-CNT sample, leading to the formation of crystal defects, which would exert a positive effect in separating the photogenerated electron–hole pairs.
It is well-acknowledged that the light absorption property of a semiconductor is intimately correlated with the energy band structure. Therefore, different photocatalysts exhibit different a light absorption property in the visible-light absorption range. Consequently, the diffuse reflectance spectra (DRS) of all samples were measured, and the results are displayed in Figure 7a. As demonstrated in the figure, the CNB sample possessed a distinct absorption edge at about 470 nm. By introducing the P element into the CNB sample, all P-CNT samples exhibited improved visible-light absorption in the visible region due to the substitution of P atoms for C to form P–N bonds [31,32]. The above results verify that the introduction of the P element can improve the light absorption of CNB, further generating more photoexcited electron–hole pairs involved in the photocatalytic reaction. In fact, the capture of visible-light could also be enhanced due to the multiple reflection effect of the incident light inside the 0.9P-CNT sample. The corresponding band-gap values from the Tauc plot (Figure 7b) decreased from 2.63 eV (CNB) to 2.55 eV (0.9P-CNT), which also indicated that such a modification strengthened the visible-light absorption. As previously reported [49], doping P atoms into g-C3N4 instead of C shifted the VBM and CBM down by 0.22 eV and 0.73 eV, respectively. These changes are consistent with previous reports that hetero-element-doped conjugated polymers have a tendency to reduce the band-gap width [50,51,52].
According to previous reports in the literature [10,35], the photocurrent response of semiconductors can effectively reflect the separation and migration of photogenerated carriers. Thus, photocurrent tests of the as-prepared samples were performed, and the results are demonstrated in Figure 8a. It was found that the photocurrents of all P-CNT samples were larger than that of CNB sample, suggesting that the introduction of the P element could accelerate the separation of photogenerated carriers and hinder the complexation between them, thus allowing more photogenerated electrons to participate in the reduction reaction. Among them, the 0.9P-CNT sample possessed the highest photocurrent response, indicating that the most efficient separation and transport of the photogenerated charges were obtained in the 0.9P-CNT sample. In addition, the separation efficiency of the interfacial charges was further explored via electrochemical impedance spectroscopy (EIS). According to previous reports [34,53], the smaller the radius of the circular arc, the lower the resistance to interfacial charge transfer from the catalyst to the reaction molecule, and the easier the separation and migration of the photogenerated charges. The EIS Nyquist plots of the CNB and all of the P-CNT samples are demonstrated in Figure 8b. Obviously, it can be seen in the figure that the 0.9P-CNT sample had the smallest radius of a circle, indicating that the as-prepared 0.9P-CNT material had the smallest internal resistance and so had more effective separation of its photogenerated charges. The results of the EIS together with the photocurrent. support that the 0.9P-CNT sample possessed the most efficient charge separation, and therefore the optimal photocatalytic hydrogen production performance can be obtained.
An energy band structure suitable for photocatalytic hydrogen production makes the P modified materials promising visible photocatalysts. The photocatalytic property was assessed by detecting the produced hydrogen under the illumination of a Xenon lamp source (λ > 420 nm). Figure 9a shows the photocatalytic hydrogen production performance of the samples, from which it could be observed that the hydrogen production of the CNB and CNT was very low. The HPR of the CNB sample was only 157.3 μmol g−1 h1, which stemmed from its narrow visible-light response range and fast-photogenerated carrier complexation rate [12,18]. The morphological modulation from the CNB to CNT (413.7 μmol g−1 h−1) increased the HPR, which illustrates that the tubular structure can improve the photocatalytic performance, but such a slight improvement is far from the desired goal. Excitingly, the HPR of all of the P-doped samples was higher than the CNB sample. With the increase in the P-source input, the HPR of P-CNT increased. When the P-source input reached 0.9 wt%, the 0.9P-CNT photocatalyst showed the highest HPR of 2749.3 μmol g−1 h−1, which was 17.5 and 6.6 times higher than the CNB and CNT samples, respectively. However, the HPR of the P-CNT gradually declined when the amount of P was further increased, indicating that too much element doping does not further enhance the catalytic performance, but causes negative effects. Furthermore, the 0.9P-CNT sample showed a better photocatalytic performance for the decomposition of aquatic hydrogen compared with most of the photocatalysts reported in the previous literature (Table 2).
The energy band structure of the 0.9P-CNT sample was studied using the Mott–Schottky (M–S) plot to ensure the band edge potential and semiconductor nature. As shown in Figure 10, the M–S plot of the 0.9P-CNT sample exhibited a positive slope, indicating a typical n-type semiconductor [55]. In addition, the flat band (fb) potential of the 0.9P-CNT sample was estimated to be −1.35 V vs. Ag/AgCl. Subsequently, its Efb (NHE) was determined using the Eq. Efb (NHE) = Efb (Ag/AgCl) + (0.059 × pH) + E(AgCl), where the pH of the Na2SO4 electrolyte was 6.8 and E(AgCl) was 0.198 V at 25 °C. Usually the ECBM = Efb (NHE) −0.2 V for n-type semiconductors, therefore, the ECBM of 0.9P-CNT was evaluated to be −0.95 V. Based on the Eg value of the Tauc plot with ECBM, the EVBM was estimated to be 1.6 V. Such a negative band can enable the 0.9P-CNT sample to achieve photocatalytic hydrogen production and become a promising hydrogen production photocatalyst.
The performance stability is very significant for the commercial use of photocatalysts, so the 0.9P-CNT sample was used for the cyclic hydrogen production test under the same conditions and the experimental results are shown in Figure 11a. After three cycles of hydrogen production tests, the amount of produced hydrogen slightly decreased. In addition, the fresh and used 0.9P-CNT samples were characterized by XRD, and the results (Figure 11b) showed that the crystal structure of the 0.9P-CNT sample remained intact after the cyclic tests, indicating that the 0.9P-CNT photocatalyst possessed excellent photostability.
To analyze the generation of unpaired electrons in the catalyst, the prepared CNB, 0.6P-CNT, 0.9P-CNT, and 1.2P-CNT samples were measured by EPR spectroscopy at room temperature, and the results are shown in Figure 12. As can be seen in Figure 12, only one Lorentz line centered at the g value of 2.0034 was observed in the magnetic field from 346.5 to 365 mT for all samples. In a typical graphitic carbon nitride structure, the g value equal to 2.0034 was attributed to the lone pair of electrons in the sp2-carbon [25]. In addition, the EPR intensity of the 0.6P-CNT, 0.9P-CNT, and 1.2P-CNT samples were all greater than that of the CNB when P atoms were introduced into the CNB. In particular, the 0.9P-CNT sample possessed the highest signal intensity, which indicates the presence of more unpaired electrons for the hydrogen production reactions in the 0.9P-CNT sample. It is also possible to show that the P element can promote the gradual development of the electronic energy band structure and inhibit the compounding of photogenerated charges. Furthermore, the above result revealed that the doping concentration of P exerted a significant effect on the generation of unpaired electrons. The optimal value for the doping concentration was 0.9 wt%, and the largest generation of unpaired electrons was achieved, which agreed well with that of photocatalytic hydrogen evolution.
Based on the analysis of the above characterization results, we propose a rational mechanism to account for the enhanced hydrogen production performance. First, a porous tube structure with a larger SSA and rough inner surface could provide more active sites for the photocatalytic reaction. Simultaneously, the presence of a mesoporous structure could accelerate the release rate of hydrogen during the hydrogen production process. Second, the doping of P into CNB led to the formation of surface defects, which acted as active sites for the photocatalytic reaction. Third, the incorporation of the P element with the optimal concentration could effectively adjust the band-gap, strengthen the harvest of visible-light, and boost the effective separation of the photogenerated charges. However, it can be noted that an excess of incorporated P atoms would evolve into recombination centers of the photogenerated electron–hole pairs, thus reducing the hydrogen production performance. Thus, we can conclude that the remarkable improvement in the photocatalytic hydrogen evolution could be due to the synergistic effect of the augmentation of active sites and light absorption capacity as well as an enhanced separation rate of the photogenerated charges.

4. Conclusions

The hollow tubular 0.9P-CNT sample was obtained by the pre-hydrothermal treatment and nitrogen calcination of melamine, and the photocatalytic HPR of the 0.9P-CNT sample could reach 2749.3 μmol g−1 h−1, which was 17.5 and 6.6 times higher than that of the CNB and CNT samples, respectively. The enhanced photocatalytic performance can be attributed to the introduction of the P elements and a porous tubular structure. The porous tubular structure allowed the 0.9P-CNT sample to have a larger SSA and rougher inner surface, which can provide more active sites for the hydrogen production reaction, therefore, the introduction of the P element can lower the band-gap, enhance the light trapping rate, and promote the effective separation of the photogenerated charges. In summary, we believe that our study not only deepens the understanding of the enchantment mechanism of photocatalytic performance, but also offers the possibility of developing stable, efficient, and noble metal-free photocatalysts for practical applications.

Author Contributions

T.J. conceived and designed the experiments; D.Y., Q.W. and K.W. performed the synthesis experiments and photocatalytic properties of the prepared samples; Z.D. and L.C. measured the prepared samples; D.Y. and T.J. analyzed the data; J.C. and P.W. participated in the discussion of the experimental data. D.Y. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Department of Science and Technology of Anhui Province China (Anhui Key Research and Development Program, 202004a05020031), the Department of Science and Technology of Henan Province, China (Henan Science and Technology Research Program, 222102520005), and the Education Department of Henan Province, China for the funding support (21A430026).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical Porous O-Doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small 2017, 13, 1603938–1603946. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, X.; Gao, D.; Yu, H.; Yu, J. High-yield lactic acid-mediated route for a g-C3N4 nanosheet photocatalyst with enhanced H2-evolution performance. Nanoscale 2019, 11, 9608–9616. [Google Scholar] [CrossRef] [PubMed]
  3. Wu, X.; Ma, H.; Zhong, W.; Fan, J.; Yu, H. Porous crystalline g-C3N4: Bifunctional NaHCO3 template-mediated synthesis and improved photocatalytic H2-evolution rate. Appl. Catal. B Environ. 2020, 271, 118899–118906. [Google Scholar] [CrossRef]
  4. Li, S.; Ng, Y.H.; Zhu, R.; Lv, S.; Wu, C.; Liu, Y.; Jing, L.; Deng, J.; Dai, H. In situ construction of elemental phosphorus nanorod-modified TiO2 photocatalysts for efficient visible-light-driven H2 generation. Appl. Catal. B Environ. 2021, 297, 120412–120419. [Google Scholar] [CrossRef]
  5. Sun, G.; Xiao, B.; Zheng, H.; Shi, J.-W.; Mao, S.; He, C.; Li, Z.; Cheng, Y. Ascorbic acid functionalized CdS-ZnO core-shell nanorods with hydrogen spillover for greatly enhanced photocatalytic H2 evolution and outstanding photostability. J. Mater. Chem. A 2021, 9, 9735–9744. [Google Scholar] [CrossRef]
  6. Dong, Y.J.; Han, Q.; Hu, Q.Y.; Xu, C.J.; Dong, C.Z.; Peng, Y.; Ding, Y.; Lan, Y.Q. Carbon quantum dots enriching molecular nickel polyoxometalate over CdS semiconductor for photocatalytic water splitting. Appl. Catal. B Environ. 2021, 293, 120214–120225. [Google Scholar] [CrossRef]
  7. Zhang, Z.; Zheng, T.; Xu, J.; Zeng, H.; Zhang, N. Carbon Quantum Dots/Bi2WO6 Composites for Efficient Photocatalytic Pollutant Degradation and Hydrogen Evolution. Nano 2017, 12, 1750082–1750090. [Google Scholar] [CrossRef] [Green Version]
  8. Dai, F.; Wang, Y.; Zhao, R.; Zhou, X.; Han, J.; Wang, L. ZnIn2S4 modified CaTiO3 nanocubes with enhanced photocatalytic hydrogen performance. Int. J. Hydrogen Energy 2020, 45, 28783–28791. [Google Scholar] [CrossRef]
  9. Xu, Y.F.; Yan, A.H.; Jiang, L.; Huang, F.; Hu, D.N.; Duan, G.H.; Zheng, F.Y. MoS2/HCSs/ZnIn2S4 nanocomposites with enhanced charge transport and photocatalytic hydrogen evolution performance. J. Alloys Compd. 2022, 895, 162504–162512. [Google Scholar] [CrossRef]
  10. Jia, T.; Fu, F.; Li, J.; Deng, Z.; Long, F.; Yu, D.; Cui, Q.; Wang, W. Rational construction of direct Z-scheme SnS/g-C3N4 hybrid photocatalyst for significant enhancement of visible-light photocatalytic activity. Appl. Surf. Sci. 2020, 499, 143941–143950. [Google Scholar] [CrossRef]
  11. Zhou, Y.; Zhang, L.; Liu, J.; Fan, X.; Wang, B.; Wang, M.; Ren, W.; Wang, J.; Li, M.; Shi, J. Brand new P-doped g-C3N4: Enhanced photocatalytic activity for H2 evolution and Rhodamine B degradation under visible light. J. Mater. Chem. A 2015, 3, 3862–3867. [Google Scholar] [CrossRef]
  12. Yu, H.; Shi, R.; Zhao, Y.; Bian, T.; Zhao, Y.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.Z.; Tung, C.H.; Zhang, T. Alkali-Assisted Synthesis of Nitrogen Deficient Graphitic Carbon Nitride with Tunable Band Structures for Efficient Visible-Light-Driven Hydrogen Evolution. Adv. Mater. 2017, 29, 1605148–1605155. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, X.; Wang, X.; Wang, F.; Yu, H. Soluble g-C3N4 nanosheets: Facile synthesis and application in photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 247, 70–77. [Google Scholar] [CrossRef]
  14. Wu, X.; Chen, F.; Wang, X.; Yu, H. In situ one-step hydrothermal synthesis of oxygen-containing groups-modified g-C3N4 for the improved photocatalytic H2-evolution performance. Appl. Surf. Sci. 2018, 427, 645–653. [Google Scholar] [CrossRef]
  15. Liang, Q.; Zhang, C.; Xu, S.; Zhou, M.; Zhou, Y.; Li, Z. In situ growth of CdS quantum dots on phosphorus-doped carbon nitride hollow tubes as active 0D/1D heterostructures for photocatalytic hydrogen evolution. J. Colloid Interf. Sci. 2020, 577, 1–11. [Google Scholar] [CrossRef] [PubMed]
  16. Zhu, Y.P.; Ren, T.Z.; Yuan, Z.Y. Mesoporous Phosphorus-Doped g-C3N4 Nanostructured Flowers with Superior Photocatalytic Hydrogen Evolution Performance. ACS Appl. Mater. Interfaces 2015, 7, 16850–16856. [Google Scholar] [CrossRef]
  17. Tian, N.; Zhang, Y.; Li, X.; Xiao, K.; Du, X.; Dong, F.; Waterhouse, G.I.N.; Zhang, T.; Huang, H. Precursor-reforming protocol to 3D mesoporous g-C3N4 established by ultrathin self-doped nanosheets for superior hydrogen evolution. Nano Energy 2017, 38, 72–81. [Google Scholar] [CrossRef]
  18. Liu, Q.; Chen, T.; Guo, Y.; Zhang, Z.; Fang, X. Ultrathin g-C3N4 nanosheets coupled with carbon nanodots as 2D/0D composites for efficient photocatalytic H2 evolution. Appl. Catal. B Environ. 2016, 193, 248–258. [Google Scholar] [CrossRef]
  19. Wang, D.; Li, J.; Xu, Z.; Zhu, Y.; Chen, G.; Cui, Z. Synthesis of g-C3N4/NiO p-n heterojunction materials with ball-flower morphology and enhanced photocatalytic performance for the removal of tetracycline and Cr6+. J. Mater. Sci. 2019, 54, 11417–11434. [Google Scholar] [CrossRef]
  20. Di, J.; Xia, J.; Yin, S.; Xu, H.; Xu, L.; Xu, Y.; He, M.; Li, H. Preparation of sphere-like g-C3N4/BiOI photocatalysts via a reactable ionic liquid for visible-light-driven photocatalytic degradation of pollutants. J. Mater. Chem. A 2014, 2, 5340–5351. [Google Scholar] [CrossRef]
  21. Jia, T.; Li, J.; Long, F.; Fu, F.; Zhao, J.; Deng, Z.; Wang, X.; Zhang, Y. Ultrathin g-C3N4 Nanosheet-Modified BiOCl Hierarchical Flower-Like Plate Heterostructure with Enhanced Photostability and Photocatalytic Performance. Crystals 2017, 7, 266. [Google Scholar] [CrossRef]
  22. Tian, Y.; Chang, B.; Fu, J.; Zhou, B.; Liu, J.; Xi, F.; Dong, X. Graphitic carbon nitride/Cu2O heterojunctions: Preparation, characterization, and enhanced photocatalytic activity under visible light. J. Solid State Chem. 2014, 212, 1–6. [Google Scholar] [CrossRef]
  23. Zhai, C.; Sun, M.; Zeng, L.; Xue, M.; Pan, J.; Du, Y.; Zhu, M. Construction of Pt/graphitic C3N4/MoS2 heterostructures on photo-enhanced electrocatalytic oxidation of small organic molecules. Appl. Catal. B Environ. 2019, 243, 283–293. [Google Scholar] [CrossRef]
  24. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation of rhodamine B and methyl orange over boron-doped g-C3N4 under visible light irradiation. Langmuir 2010, 26, 3894–3901. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, G.; Zhang, M.; Ye, X.; Qiu, X.; Lin, S.; Wang, X. Iodine modified carbon nitride semiconductors as visible light photocatalysts for hydrogen evolution. Adv. Mater. 2014, 26, 805–809. [Google Scholar] [CrossRef] [PubMed]
  26. Li, J.; Shen, B.; Hong, Z.; Lin, B.; Gao, B.; Chen, Y. A facile approach to synthesize novel oxygen-doped g-C3N4 with superior visible-light photoreactivity. Chem. Commun. 2012, 48, 12017–12019. [Google Scholar] [CrossRef]
  27. Dong, G.; Zhao, K.; Zhang, L. Carbon self-doping induced high electronic conductivity and photoreactivity of g-C3N4. Chem. Commun. 2012, 48, 6178–6180. [Google Scholar] [CrossRef]
  28. Ma, X.; Lv, Y.; Xu, J.; Liu, Y.; Zhang, R.; Zhu, Y. A Strategy of Enhancing the Photoactivity of g-C3N4 via Doping of Nonmetal Elements: A First-Principles Study. J. Phys. Chem. C 2012, 116, 23485–23493. [Google Scholar] [CrossRef]
  29. Jiang, L.; Yuan, X.; Zeng, G.; Chen, X.; Wu, Z.; Liang, J.; Zhang, J.; Wang, H.; Wang, H. Phosphorus-and Sulfur-Codoped g-C3N4: Facile Preparation, Mechanism Insight, and Application as Efficient Photocatalyst for Tetracycline and Methyl Orange Degradation under Visible Light Irradiation. ACS Sustain. Chem. Eng. 2017, 5, 5831–5841. [Google Scholar] [CrossRef]
  30. Ran, J.; Ma, T.Y.; Gao, G.; Du, X.-W.; Qiao, S.Z. Porous P-doped graphitic carbon nitride nanosheets for synergistically enhanced visible-light photocatalytic H2 production. Energy Environ. Sci. 2015, 8, 3708–3717. [Google Scholar] [CrossRef]
  31. Wu, M.; Zhang, J.; He, B.-b.; Wang, H.-w.; Wang, R.; Gong, Y.-s. In-situ construction of coral-like porous P-doped g-C3N4 tubes with hybrid 1D/2D architecture and high efficient photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2019, 241, 159–166. [Google Scholar] [CrossRef]
  32. Guo, S.; Tang, Y.; Xie, Y.; Tian, C.; Feng, Q.; Zhou, W.; Jiang, B. P-doped tubular g-C3N4 with surface carbon defects: Universal synthesis and enhanced visible-light photocatalytic hydrogen production. Appl. Catal. B Environ. 2017, 218, 664–671. [Google Scholar] [CrossRef]
  33. Putri, L.K.; Ng, B.-J.; Er, C.-C.; Ong, W.-J.; Chang, W.S.; Mohamed, A.R.; Chai, S.-P. Insights on the impact of doping levels in oxygen-doped gC3N4 and its effects on photocatalytic activity. Appl. Surf. Sci. 2020, 504, 144427–144437. [Google Scholar] [CrossRef]
  34. Mo, Z.; Xu, H.; Chen, Z.; She, X.; Song, Y.; Wu, J.; Yan, P.; Xu, L.; Lei, Y.; Yuan, S.; et al. Self-assembled synthesis of defect-engineered graphitic carbon nitride nanotubes for efficient conversion of solar energy. Appl. Catal. B Environ. 2018, 225, 154–161. [Google Scholar] [CrossRef]
  35. Wang, D.; Li, Y.; Yu, B.; Li, H.; Jiang, W.; Deng, X.; Wen, Y.; Liu, C.; Che, G. Improved visible-light driven photocatalysis by loading Au onto C3N4 nanorods for degradation of RhB and reduction of CO2. Adv. Powder Technol. 2021, 32, 1653–1662. [Google Scholar] [CrossRef]
  36. Zhang, X.; Zhang, Y.; Jia, X.; Zhang, N.; Xia, R.; Zhang, X.; Wang, Z.; Yu, M. In situ fabrication of a novel S-scheme heterojunction photocatalyts Bi2O3/P-C3N4 to enhance levofloxacin removal from water. Sep. Purif. Technol. 2021, 268, 118691–118702. [Google Scholar] [CrossRef]
  37. Wang, Y.; Zheng, M.; Zhao, H.; Qin, H.; Fan, W.; Zhao, X. Zn2GeO4-x/ZnS heterojunctions fabricated via in situ etching sulfurization for Pt-free photocatalytic hydrogen evolution: Interface roughness and defect engineering. Phys. Chem. Chem. Phys. 2020, 22, 10265–10277. [Google Scholar] [CrossRef]
  38. Yin, M.; Wu, C.; Jia, F.; Wang, L.; Zheng, P.; Fan, Y. Efficient photocatalytic hydrogen production over eosin Y-sensitized MoS2. RSC Adv. 2016, 6, 75618–75625. [Google Scholar] [CrossRef]
  39. Che, H.; Che, G.; Zhou, P.; Liu, C.; Dong, H. Yeast-derived carbon sphere as a bridge of charge carriers towards to enhanced photocatalytic activity of 2D/2D Cu2WS4/g-C3N4 heterojunction. J. Colloid Interf. Sci. 2019, 546, 262–275. [Google Scholar] [CrossRef]
  40. Huo, T.; Ba, G.; Deng, Q.; Yu, F.; Wang, G.; Li, H.; Hou, W. A dual strategy for synthesizing carbon/defect comodified polymeric carbon nitride porous nanotubes with boosted photocatalytic hydrogen evolution and synchronous contaminant degradation. Appl. Catal. B Environ. 2021, 287, 119995–120006. [Google Scholar] [CrossRef]
  41. Lee, H.; Kumbhar, V.S.; Lee, J.; Oh, H.; Lee, K. Boosted photocatalytic hydrogen evolution by tuning inner pore size and co-catalyst thickness of the anodic TiO2 nanotubes. Catal. Today 2021, 359, 3–8. [Google Scholar] [CrossRef]
  42. Wu, W.; Song, Y.; Zhou, P.; Yu, Y.; Xie, L.; Hua, L. Mechanical properties improvement of pre-deformed Al–Zn–Mg–Cu alloys by electroshocking treatment based on the non-equilibrium scattering of electron-dislocation. J. Alloys Compd. 2021, 861, 157987–157994. [Google Scholar] [CrossRef]
  43. Chen, L.; Wang, Y.; Cheng, S.; Zhao, X. Nitrogen defects/boron dopants engineered tubular carbon nitride for efficient tetracycline hydrochloride photodegradation and hydrogen evolution. Appl. Catal. B Environ. 2022, 33, 120932–120941. [Google Scholar] [CrossRef]
  44. Dementjev, A.P.; Graaf, A.d.; Van de Sanden, M.C.M.; Maslakov, K.I.; Naumkin, A.V.; Serov, A.A. X-Ray photoelectron spectroscopy reference data for identification of the C N phase in carbon-nitrogen films. Diam. Relat. Mater. 2000, 9, 1904–1907. [Google Scholar] [CrossRef]
  45. Zhang, X.; Xie, X.; Wang, H.; Zhang, J.; Pan, B.; Xie, Y. Enhanced photoresponsive ultrathin graphitic-phase C3N4 nanosheets for bioimaging. J. Am. Chem. Soc. 2013, 135, 18–21. [Google Scholar] [CrossRef]
  46. Li, X.H.; Xu, H.Y.; Zhang, X.T.; Liu, Y.C.; Sun, J.W.; Lu, Y.M. Local chemical states and thermal stabilities of nitrogen dopants in ZnO film studied by temperature-dependent x-ray photoelectron spectroscopy. Appl. Phys. Lett. 2009, 95, 191903–191906. [Google Scholar] [CrossRef]
  47. Li, K.; Zeng, Z.; Yan, L.; Luo, S.; Luo, X.; Huo, M.; Guo, Y. Fabrication of platinum-deposited carbon nitride nanotubes by a one-step solvothermal treatment strategy and their efficient visible-light photocatalytic activity. Appl. Catal. B Environ. 2015, 165, 428–437. [Google Scholar] [CrossRef]
  48. Chen, W.; Liu, T.Y.; Huang, T.; Liu, X.H.; Yang, X.J. Novel mesoporous P-doped graphitic carbon nitride nanosheets coupled with ZnIn2S4 nanosheets as efficient visible light driven heterostructures with remarkably enhanced photo-reduction activity. Nanoscale 2016, 8, 3711–3719. [Google Scholar] [CrossRef]
  49. Liu, J.J. Effect of phosphorus doping on electronic structure and photocatalytic performance of g-C3N4: Insights from hybrid density functional calculation. J. Alloys Compd. 2016, 672, 271–276. [Google Scholar] [CrossRef]
  50. Hu, S.; Ma, L.; You, J.; Li, F.; Fan, Z.; Wang, F.; Liu, D.; Gui, J. A simple and efficient method to prepare a phosphorus modified g-C3N4 visible light photocatalyst. RSC Adv. 2014, 4, 21657–21663. [Google Scholar] [CrossRef]
  51. Guo, S.; Deng, Z.; Li, M.; Jiang, B.; Tian, C.; Pan, Q.; Fu, H. Phosphorus-Doped Carbon Nitride Tubes with a Layered Micro-nanostructure for Enhanced Visible-Light Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2016, 55, 1830–1834. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, Y.Z.; Chen, Z.W.; Li, J.L.; Lu, Z.Y.; Wang, X. Self-assembled synthesis of oxygen-doped g-C3N4 nanotubes in enhancement of visible-light photocatalytic hydrogen. J. Energy Chem. 2021, 54, 36–44. [Google Scholar] [CrossRef]
  53. Jiang, W.; Zong, X.; An, L.; Hua, S.; Miao, X.; Luan, S.; Wen, Y.; Tao, F.F.; Sun, Z. Consciously Constructing Heterojunction or Direct Z-Scheme Photocatalysts by Regulating Electron Flow Direction. ACS Catal. 2018, 8, 2209–2217. [Google Scholar] [CrossRef]
  54. Huang, Z.; Li, F.; Chen, B.; Yuan, G. Hydrogen from Water over Openly-Structured Graphitic Carbon Nitride Polymer through Photocatalysis. ChemSusChem 2016, 9, 478–484. [Google Scholar] [CrossRef]
  55. Liu, B.; Qiao, M.; Wang, Y.; Wang, L.; Gong, Y.; Guo, T.; Zhao, X. Persulfate enhanced photocatalytic degradation of bisphenol A by g-C3N4 nanosheets under visible light irradiation. Chemosphere 2017, 189, 115–122. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic illustration for the synthesis of P-CNT.
Scheme 1. Schematic illustration for the synthesis of P-CNT.
Nanomaterials 12 01759 sch001
Figure 1. The XRD patterns of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Figure 1. The XRD patterns of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Nanomaterials 12 01759 g001
Figure 2. The SEM images of the CNB (a), low magnification (b), and high magnification (c) of the supramolecular precursors, and low magnification (d) and high magnification (e,f) of the 0.9P-CNT sample.
Figure 2. The SEM images of the CNB (a), low magnification (b), and high magnification (c) of the supramolecular precursors, and low magnification (d) and high magnification (e,f) of the 0.9P-CNT sample.
Nanomaterials 12 01759 g002
Figure 3. The TEM images (a,b) and HRTEM images (c,d) of the 0.9P-CNT sample, the inset is the corresponding selected area electron diffraction image, HAADF-STEM image and element mapping images of C, N, and P for the 0.9P-CNT sample (eh), EDX spectrum of the 0.9P-CNT sample (i).
Figure 3. The TEM images (a,b) and HRTEM images (c,d) of the 0.9P-CNT sample, the inset is the corresponding selected area electron diffraction image, HAADF-STEM image and element mapping images of C, N, and P for the 0.9P-CNT sample (eh), EDX spectrum of the 0.9P-CNT sample (i).
Nanomaterials 12 01759 g003
Figure 4. (a) ( 2 ¯ 10 ), (b) ( 1 ¯ 20 ), (c) (110) inverse Fourier transform images in the single crystal region.
Figure 4. (a) ( 2 ¯ 10 ), (b) ( 1 ¯ 20 ), (c) (110) inverse Fourier transform images in the single crystal region.
Nanomaterials 12 01759 g004
Figure 5. N2 adsorption-desorption curves of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Figure 5. N2 adsorption-desorption curves of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Nanomaterials 12 01759 g005
Figure 6. The XPS survey spectra of the as-prepared CNB and 0.9P-CNT samples (a), and high-resolution XPS spectra of CNB and 0.9P-CNT in different regions. (b) C1s, (c) N1s, and (d) P2p.
Figure 6. The XPS survey spectra of the as-prepared CNB and 0.9P-CNT samples (a), and high-resolution XPS spectra of CNB and 0.9P-CNT in different regions. (b) C1s, (c) N1s, and (d) P2p.
Nanomaterials 12 01759 g006
Figure 7. The UV–Vis diffuse reflectance spectra (a) and corresponding Tauc plot (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Figure 7. The UV–Vis diffuse reflectance spectra (a) and corresponding Tauc plot (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Nanomaterials 12 01759 g007
Figure 8. The photocurrent (a) and electrochemical impedance spectra (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Figure 8. The photocurrent (a) and electrochemical impedance spectra (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT samples.
Nanomaterials 12 01759 g008
Figure 9. The photocatalytic hydrogen production (a) and HPR (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, 1.5P-CNT, and CNT samples.
Figure 9. The photocatalytic hydrogen production (a) and HPR (b) of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, 1.5P-CNT, and CNT samples.
Nanomaterials 12 01759 g009
Figure 10. The Mott–Schottky plot of the 0.9P-CNT sample.
Figure 10. The Mott–Schottky plot of the 0.9P-CNT sample.
Nanomaterials 12 01759 g010
Figure 11. The cyclic hydrogen production test of the 0.9P-CNT sample (a) and the XRD patterns of the fresh and used 0.9P-CNT samples (b).
Figure 11. The cyclic hydrogen production test of the 0.9P-CNT sample (a) and the XRD patterns of the fresh and used 0.9P-CNT samples (b).
Nanomaterials 12 01759 g011
Figure 12. Room temperature EPR signals for the CNB, 0.6P-CNT, 0.9P-CNT, and 1.2P-CNT samples.
Figure 12. Room temperature EPR signals for the CNB, 0.6P-CNT, 0.9P-CNT, and 1.2P-CNT samples.
Nanomaterials 12 01759 g012
Table 1. The SSA of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT.
Table 1. The SSA of the CNB, 0.3P-CNT, 0.6P-CNT, 0.9P-CNT, 1.2P-CNT, and 1.5P-CNT.
SamplesSSA (m2 g−1)
CNB9.4
0.3P-CNT23.2
0.6P-CNT25
0.9P-CNT31.5
1.2P-CNT26.3
1.5P-CNT16.5
Table 2. The photocatalytic hydrogen production performance on the various reported photocatalysts.
Table 2. The photocatalytic hydrogen production performance on the various reported photocatalysts.
PhotocatalystsPtHPR (μmol g−1 h−1)Light Source (>420 nm)References
porous crystalline g-C3N4 1 wt%1010LED (12 W)[3]
CdS/P-g-C3N4 tube-1%1 wt%1579300 W Xenon lamp[15]
S-g-C3N4/g-C3N4 (0.3%)1 wt%359.6LED (12 W)[13]
g-C3N41 wt%646LED (12 W)[2]
P-g-C3N4 tube3 wt%2020300W Xenon lamp[31]
P-TCN1 wt%570300W Xenon lamp[32]
P-g-C3N4-5503 wt%506300 W Xenon lamp[11]
P-g-C3N4 nanosheet1 wt%1596300 W Xenon lamp[30]
CNMS-6003 wt%3920300 W Xenon lamp[54]
0.9P-CNT1 wt%2749.3300 W Xenon lampThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, D.; Jia, T.; Deng, Z.; Wei, Q.; Wang, K.; Chen, L.; Wang, P.; Cui, J. One-Dimensional P-Doped Graphitic Carbon Nitride Tube: Facile Synthesis, Effect of Doping Concentration, and Enhanced Mechanism for Photocatalytic Hydrogen Evolution. Nanomaterials 2022, 12, 1759. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12101759

AMA Style

Yu D, Jia T, Deng Z, Wei Q, Wang K, Chen L, Wang P, Cui J. One-Dimensional P-Doped Graphitic Carbon Nitride Tube: Facile Synthesis, Effect of Doping Concentration, and Enhanced Mechanism for Photocatalytic Hydrogen Evolution. Nanomaterials. 2022; 12(10):1759. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12101759

Chicago/Turabian Style

Yu, Dazhuang, Tiekun Jia, Zhao Deng, Qichen Wei, Kun Wang, Lihua Chen, Pingping Wang, and Jiedong Cui. 2022. "One-Dimensional P-Doped Graphitic Carbon Nitride Tube: Facile Synthesis, Effect of Doping Concentration, and Enhanced Mechanism for Photocatalytic Hydrogen Evolution" Nanomaterials 12, no. 10: 1759. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12101759

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop