Broadening the light absorption range and suppressing the carrier complexation are the two keys to enhance the photocatalytic activity. In this work, a novel two-dimensional (2D) photocatalyst was successfully prepared by modified hydrothermal method and applied in tetracycline (TC) degradation. The degradation rate of CD(Cu)-Ni-MOL for TC reached 93.5% within 60 min under the visible light condition. The improved photocatalytic performance of CD(Cu)-Ni-MOL was attributed to the constructed 2D layered structure and the special properties of CD(Cu). The doped Cu in carbon dots (CDs) exhibited excellent photocatalytic performance among the elements of Cu, Zn, Ni, Co and Fe. The order of photocatalytic performance improvement was Cu > Zn > Ni > Co > Fe. In addition, a possible degradation pathway for TC was proposed. This work confirms the great potential of CD(Cu)-Ni-MOL as a highly efficient photocatalyst in removing tetracycline pollutants in water.

  • A novel 2D photocatalyst of single atom-dispersed copper and carbon quantum dots co-loaded with ultrathin Ni-MOL was firstly prepared.

  • The incorporation of CD(Cu) induced electronic structure change of CD(Cu)-Ni-MOL were analyzed.

  • The influence of Cu dopants on photocatalytic performance are discussed.

Recently, chemical pharmaceuticals (antibiotics) received more and more attention due to their negative effects and possible risks to ecosystems (Li et al. 2018). Tetracycline (TC), a typical category of sulfonamide antimicrobial agents, is applied widely in the field of human medicine. The environmental concentration is relatively low (ng/L-mg/L). Some techniques have been used to eliminate TC from water, including adsorption, liquid membrane, photocatalysis and so on (Barhoumi et al. 2016). Photocatalysis is considered to be one of the most progressive methods to remove TC because it is environmentally friendly, and highly stable and cheap. Many photocatalysts have been used to degrade antibiotics from water, including g-C3N4, TiO2 and CdS. However, most of those photocatalysts are facing two challenges in application: they can only absorb ultraviolet region photoenergy and the rate of recombination of photogenerated holes with electrons is high. Thus, designing a photocatalyst with high efficiency of photoelectron-hole separation ability and effectively harvesting visible light has attracted more and more attention in the last few decades (Li et al. 2011).

Owing to low bandgap energy and excellent photo energy harvesting efficiency, organic nanostructured semiconductors have attracted more and more attention. Large numbers of organic-based semiconductor photocatalysis have been constructed with conjugated porphyrins as institutional building blocks to build solid-state soft materials. Metalloporphyrin have garnered special attention due to several important characteristics for photocatalysis, including higher triplet state quantum yields, long-lived excited states, and intense visible absorption (Zhu et al. 2020). The heterogenization of functional groups with long-range-ordered structures in crystalline frameworks will reflect different photocatalytic properties when porphyrins are used as functional molecules in homogeneous systems. Micheroni et al. have constructed porphyrin metal-organic frameworks (MOF) using of 5,15-bis(4-carboxyphenyl) porphyrin (H2BCPP) porphyrin linkers (Micheroni et al. 2018). Johnson et al. reported that MOFs with octatopic linkers and tetratopic linkers porphyrin possessed a distinct structure (Johnson et al. 2014). Most of those studies focused on the construction of large conjugate system, however, few studies focused on derivatives of porphyrin, such as based-porphyrin carbon quantum dots. Carbon dots (CDs) is an innovative carbon nanomaterial with several advantages, such as good electron conductivity, low toxicity, adjustable fluorescent emission, low cost and facile functionalization. Especially, the up-conversion luminescence characteristic of CDs permits it to directly capture the near-infrared portion of solar energy and convert it to visible light. Compared with traditional carbon source CDs, based-porphyrin CDs possess some intrinsic advantages, such as high N/C atomic ratios, low carbonization temperature and low melting point (Wang et al. 2020a, 2020b). More importantly, it can give rise to metal atomically dispersed in CDs. In recent years, metal-doping have emerged as a new hotspot of activity in photocatalysis with the highest catalytic efficiency. Moreover, as a result of altering electronic and energy level structure, the unique activity of the ligand atoms may be enhanced, resulting in further adjustment in adsorbing visible photoenergy. Compared with other transition metal, copper (Cu) can be easily reduced to zero value and is often used as a substitute for precious metals. In addition, Cu is a variable metal with +1 and +2 valence, which is favorable for the transfer of photocarriers in photocatalysts. However, to our knowledge, the combination of Cu and based-porphyrin CDs has not been investigated yet.

However, it is not convenient to separate CDs from solution due to the characteristics of high dispersion and high stability. Hence, supporting CDs on solid materials is very essential. Recently, the combination of semiconductor materials and MOFs has become a new research trend, which has three significant advantages. Firstly, it avoids the compounding of photo-generated carries by forming Schottky heterojunctions between MOFs semiconductors and precious metal centers. Secondly, it can absorb more visible light by effectively reducing the energy band gap. It is possible to form surface plasmon resonance (SPR) on contact surfaces between the MOF and the semiconductor material. Thirdly, with a high specific surface area and numerous reaction sites, MOFs can hasten the accumulation of pollutants on the composite surface and increase the degradation rate (Pelaeza et al. 2012). Therefore, loading CDs onto MOFs could dramatically enhance the photocatalytic efficiency via two pathways. The first way is that CDs can accelerate the separation of photogenerated electron-holes due to their good electro-conductivity. The second one is that, due to their up-conversion luminescence, CDs can convert the near-infrared light to visible light, thus the light utilization efficiency can be greatly enhanced (Huang et al. 2017). In addition, it is an ideal support for CDs on MOFs with plentiful active spots and extremely large area of specific surface. Especially, two-dimensional (2D) MOFs possess more active sites and loading capacity, in comparison with three-dimensional (3D) MOFs. However, to our knowledge, constructing 2D MOFs supporting CDs to enhance the photocatalytic performance has not been reported.

Herein, in this work, a hybrid photocatalyst combining Ni-MOL (one of MOFs, Ni-metal organic layers) and CDs derived from porphyrin was prepared for the first time. Characterization of the physico-chemical performance of the prepared materials was conducted. In addition, the charge transfer and photodegradation mechanism of TC nanocomposites were discussed.

Materials

Para-phtalic acid (PTA, ≥99%), pyrrole (C4H5N, AR), methyl 4-formylbenzoate (C9H8O3, ≥98%), formic acid (CH2O2, ≥85%), potassium hydroxide (KOH, 90%), potassium dichromate (≥99.8%), and sodium hydroxide (NaOH, ≥96.0%) were purchased from Mackiln Chemical Co., Ltd Propionic acid (C3H6O2, ≥99.5%), tetrahydrofuran (C4H8O, ≥99.5%), methanol (CH4O, ≥99.5%), copper(II) chloride dihydrate (CuCl2·2H2O, ≥99.0%), isopropyl alcohol (IPA, ≥99.5%), N,N-dimethylformamide (DMF, ≥99.5%), para-benzoquinone (p-BQ, ≥99.0%), nickel(II) nitrate hexahydrate (Ni(NO3)2·6H2O, ≥98.0%), and ethylenediaminetetraacetic acid disodium (EDTA-2Na, ≥98.0%) were purchased from Shanghai Test. HPLC-methanol (≥99.999%) were bought from Kermel. Hydrochloric acid (HCl, 36.0–38.0%) and diethyl ether anhydrous ((CH3CH2)2O, ≥99.6%) were purchased from Yantai Yuandong Chemicals Co., Ltd.

Synthesis of CD(Cu)-Ni-MOL

Preparation of Ni-MOL

The method used to prepare Ni-MOL can be found in literature (Yang et al. 2021). PTA (0.166 g) was dissolved into 5 mL DMF solution with stirring for 10 min to form solution A. Ni(NO3)2⋅6H2O (0.436 g) was added into 10 mL DMF solution with stirring for 10 min to form solution B. Subsequently, solution A was added drop-wise into solution B. Afterwards, 2 mL water containing NaOH (3.2 mg, 0.08 mmol) was slowly added into the mixed solution and stirring for 1 h. Thereafter, the mixture was transferred into Teflon-lined stainless-steel autoclave and heated at 120 °C for 10 h. The resulting precipitate was collected by filtration and washed with DMF and alcohol for several times. The collected sample was dried at 50 °C for 12 h in an oven to obtain the final product.

Preparation of tetracharboxylic methyl porphyrin

Pyrrole (3.0 g, 0.043 mol) and methyl 4-formylbenzoate (6.9 g, 0.042 mol) were placed in a three-necked flask. The mixtures were completely dissolved by 100 mL propionic acid. Subsequently, the mixture was placed in an oil bath at 160 °C for 12 h, followed by cooling down. The solid material was collected after filtration and washed with ethanol, ether and tetrahydrofuran. After drying, a purple substance was obtained.

Preparation of tetracharboxylic porphyrin (TCPP)

Tetracarboxymethyl porphyrin (0.99 g) was dissolved in 30 mL tetrahydrofuran and 30 mL methanol. Then, 30 mL potassium hydroxide solution (containing 6.82 g potassium hydroxide) was added into the mixed solution, followed by being refluxed in an oil bath at 80 °C for 12 h. The solids were acidified with 1M acid solution, filtered and washed with water afterwards. The final product (TCPP) was dried in an oven.

Preparation of CD(Cu)-Ni-MOL

TCPP was impregnated in 0.5 mol/L CuCl2•2H2O solution and sonicated for 30 min. Later, it was settled down for 24 h. After soaking, 1 g of the soaked TCPP was calcined at 250 °C for 2 h with the rate of 5 °C/min in N2. Later, the material and NaOH solution (100 ml, 0.15 mol/L) were added into a Teflon autoclave at 80 °C for 24 h. Following alkali etching, the CD(Cu) suspension was centrifuged at 10,000 rpm/min for 5 min. Subsequently, Ni-MOL (0.2 g) was added into 50 mL of supernatant and hydrothermal reaction was carried out at 90 °C for 3 h. Then, the precipitate was washed with ultrapure water.

Crystal structure and surface morphology

Crystal structure (TEM and HRTEM) of CD(Cu), Ni-MOL and CD(Cu)-Ni-MOL, EDS mapping image of CD(Cu)-Ni-MOL and AFM of Ni-MOL and CD(Cu)-Ni-MOL are shown in Figure 1. CD(Cu) with an average diameter of about 5–10 nm was a circular carbon nanodots (Figure 1(a)). The microscopic morphology of Ni-MOL was 2D layered construction (Figure 1(b)). Figure 1(c) shows that CD(Cu)-Ni-MOL was also a 2D layered structure and a large number of CDs had loaded on the surface of Ni-MOL. Crystal lattice stripes (0.31 and 0.36 nm, Figure 1(d)) were corresponding to Ni-MOL (201) and CD(Cu) (110), respectively. In addition, the energy dispersive spectroscopy (EDS) mapping proved that the C, N, O, Cu and Ni elements were uniformly distributed (Figure 1(e)), suggesting that the synthesis of CD(Cu) and Ni-MOL was successful. Figure 1(f) shows the atomic force microscope (AFM) images of Ni-MOL and CD(Cu)-Ni-MOL. It can be found that the surface of Ni-MOL was smooth (Figure 1(f)). After loading CD(Cu), the surface of CD(Cu)-Ni-MOL became rough and the irregular sphere particles could be clearly observed, indicating that CD(Cu) was distributed on the outer surface of Ni-MOL. It has been reported that the smoother surface would result in a weaker light absorption and lower photocatalytic activity (Li et al. 2020). Hence, the rough surface of CD(Cu)-Ni-MOL is conducive to the improvement of its catalytic performance.
Figure 1

TEM and HRTEM images of CD(Cu) (a), Ni-MOL (b), CD(Cu)-Ni-MOL (c, d), EDS mapping image of CD(Cu)-Ni-MOL (e) and AFM of Ni-MOL (f) and CD(Cu)-Ni-MOL (g).

Figure 1

TEM and HRTEM images of CD(Cu) (a), Ni-MOL (b), CD(Cu)-Ni-MOL (c, d), EDS mapping image of CD(Cu)-Ni-MOL (e) and AFM of Ni-MOL (f) and CD(Cu)-Ni-MOL (g).

Close modal
X-ray diffraction (XRD) pattern of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL are shown in Figure 2(a). As shown in Figure 2(a), peaks at 16.9° and 8.5° were corresponding to (010) and (100) crystalline planes of Ni-MOL phase structure. For CDs, the diffraction peak at 2θ = 8.1°, 22.9° and 31.6° were derived from the interlayer reflection of a graphite-like CDs crystal structure. After the introduction of Cu in CDs, the peak pattern became blurred and the peak located at 31.6° disappeared, suggesting that the addition of Cu had broken graphite-like layers of the CDs. In another words, Cu has been successfully doped in CDs. All of the characteristic peaks of Ni-MOL and CDs were founded in CD-Ni-MOL. In addition, new peaks located at 6.7° and 7.6° were found in CD-Ni-MOL, which corresponded to deca-dodecasil-3R(PDF:41-0571). This indicates that part of Ni-MOL was decomposed during the hydrothermal preparation of CD-Ni-MOL. By contrast with CD-Ni-MOL, the intensity peaks at 22.9° further reduced in CD(Cu)-Ni-MOL, indicating that CD(Cu) have loaded on CD(Cu)-Ni-MOL.
Figure 2

(a) X-ray diffraction (XRD) pattern of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL; (b) FT-IR spectrogram of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL.

Figure 2

(a) X-ray diffraction (XRD) pattern of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL; (b) FT-IR spectrogram of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL.

Close modal

The FT-IR spectra of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL are presented in Figure 2(b). Peaks sited at 3,433 cm−1 were accorded to the stretching vibrations of O-H of samples. For Ni-MOL, a series of peaks located at 1,446–1,680 cm−1 were related to the feature peaks of benzene rings. A peak appearing at 823 cm−1 belonged to the feature peak of Ni-O (Yang et al. 2021). A peak appearing at 1,750 cm−1 was ascribed to the carbonyl group of CD and CD(Cu). The characteristic peak of C-N of CDs and CD(Cu) was located at 1,350 cm−1. Compared with CD, the new peak presented at 1,004 cm−1 was ascribed to Cu-O in CD(Cu). All characteristic peaks of CD and Ni-MOL were detected in the CD-Ni-MOL and CD(Cu)-Ni-MOL, proving the successful combination of CD and Ni-MOL. The absorption peak of C = O of CDs was blue shifted to 1,700 cm−1 due to the influence of the Ni as the electron withdrawing group and the combination with Ni-MOL molecular bond, which was covered by the characteristic peaks of benzene ring.

Figure 3 shows the TGA results of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL. Thermo gravimetric curves of Ni-MOL could be devided as two decomposition stages. The first stage with a loss of 8.22% from 160 °C to 289 °C was corresponding to the volatilization of the adsorbed solvent and crystallization water. The pyrolysis of benzene ring corresponded to the second stage with a loss of 36.46% from 362 °C to 452 °C. However, there were three decomposition stages for CDs. The first stage with a loss of 13.5% from 40 °C to 280 °C was ascribed to the volatilization of adsorbed water (Wang et al. 2020a, 2020b). The second stage was accorded to the decomposition of pyrrole, carbonyl and carboxyl. The third stage was accorded to the decomposition of hydroxyl. It is interesting that the addition of Cu could accelerate the decomposition of CD at the range of 50–180 °C. This may be due to the fact that Cu as catalyst accelerated the splitting of porphyrin ring. With respect to the CD-Ni-MOL and CD(Cu)-Ni-MOL, the pyrolysis curves were similar to Ni-MOL except at the range of 400–450 °C. This was due to the low pyrolysis loss of CD in this region. The phenomenon that the addition of Cu leaded to intensified pyrolysis loss also appeared in the comparison between CD-Ni-MOL and CD(Cu)-Ni-MOL. Different pyrolysis losses were mainly reflected in two regions. The first region located at 200–300 °C was corresponding to the pyrolysis of porphyrin ring. The second region located at 470 °C was ascribed to the pyrolysis of carbonyl, indicating that the addition of Cu had changed the ratio of carbonyl.
Figure 3

Thermal gravity analysis (TGA) results of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL.

Figure 3

Thermal gravity analysis (TGA) results of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL.

Close modal
The XPS spectra of CD(Cu), Ni-MOL and CD(Cu)-Ni-MOL were illustrated in Figure 4. The peaks of all elements of CD(Cu) and Ni-MOL had been found in CD(Cu)-Ni-MOL. C 1s spectrum (Figure 4(b)) of CD(Cu), Ni-MOL and CD(Cu)-Ni-MOL located at 284.13 and 287.81 eV could be conformed to two different carbon states, assigning to various carbon C-C(sp2) and C = C-O. The N 1s spectrum showed the peaks at 397.78 and 399.85 eV, of which were ascribed to the binding energy of amino and imine, respectively. Compared with CD(Cu), the peak area of imine increased, while it was decreased for amino, suggesting that the synthesis of Ni-MOL and CD(Cu) had changed the chemical environment of N atom. The spectrum of O 1s was exhibited in Figure 4(d). Peaks located at 535.37 eV and 530.97 eV can be ascribed to C-O-Na and Cu–O bonds in CD(Cu), respectively. As for Ni-MOL, peaks at 530.84 eV and 532.08 eV were corresponding to the Ni-O and benzene carbonyl, respectively. As for CD(Cu)-Ni-MOL, 530.93 eV was ascribed to Cu-O and Ni-O, and 532.55 eV was corresponding to benzene carbonyl. Compared with Ni-MOL, the characteristic peak of benzene carbonyl has blue shifted 0.47 eV, which may due to the fact that the benzene carbonyl had reacted with C-O-Na. This could explain the disappearance of peak C-O-Na. The XPS of Cu 2p spectrum (Figure 4(e)) exhibited two fitted peaks, which were belonged to Cu 2p3/2 and Cu 2p1/2 groups. Two obvious peaks were consisted with Cu0 and Cu2+ respectively. The spectrum of Ni was exhibited in Figure 4(f). Peaks appeared at 879.18, 873.26, 861.03, and 855.59 eV were assigned to Ni 2p1/2 sat, Ni 2p1/2, Ni 2p3/2 sat and Ni 2p3/2, respectively. The spin–orbit separations of Ni was 17.09 eV, indicating that the Ni was appeared as Ni2+ in CD(Cu)-Ni-MOL (Saranya & Selladurai 2019).
Figure 4

XPS spectra of CD(Cu) and CD(Cu)-Ni-MOL (a) survey specturm, (b) C 1s, (c) N 1s, (d) O 1s, (e) Cu 2p and (f) Ni 2p.

Figure 4

XPS spectra of CD(Cu) and CD(Cu)-Ni-MOL (a) survey specturm, (b) C 1s, (c) N 1s, (d) O 1s, (e) Cu 2p and (f) Ni 2p.

Close modal

Material optical properties

The band gap energy of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL were 3.536 eV, 2.724 eV, 2.677 eV, 2.657 eV and 2.071 eV, respectively (Figure 5(a) inset). The Ni-MOL loaded with CD could decrease the band gap, due to the fact that CD could be stimulated partially by visible and near-infrared light. Low-energy visible light could be converted to high-energy visible light through the upconversion effect. The luminescence of CD was utilized by coupling of Ni-MOL with CD, leading to the broadening of light absorption range and the improvement in photocatalytic performance. It was interesting that the presence of Cu could alter the band gap. Compared with CD and CD-Ni-MOL, the band gap of CD(Cu) and CD(Cu)-Ni-MOL had reduced by 0.047 eV and 0.586 eV, respectively. It suggested that the incorporation of Cu could enhance the detachment of the photogenerated electrons and holes. In addition, the transformation from Cu2+ to Cu0 could greatly decrease the band gap (Selvakumar et al. 2022).
Figure 5

(a) UV-vis diffuse reflectance spectra (DRS) of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL. The inset shows optical band gaps of prepared materials; (b) photoluminescence (PL) spectra of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL.

Figure 5

(a) UV-vis diffuse reflectance spectra (DRS) of CD, CD(Cu), Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL. The inset shows optical band gaps of prepared materials; (b) photoluminescence (PL) spectra of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL.

Close modal

Figure 5(b) shows the photoluminescence (PL) spectra of Ni-MOL, CD, CD(Cu), CD-Ni-MOL and CD(Cu)-Ni-MOL. The fluorescence intensities were different at 720 nm for all samples. In general, a higher intensity of PL emission resulted in a higher rate of photogenerated electron/hole pair complexation. CD(Cu)-Ni-MOL emission peak possessed the lowest peak intensity among all the samples, which attributed to the valid charge transfer between Ni-MOL and CD particles. Thus, it decreased the carrier complication process. In addition, CD(Cu) had lower PL emission intensity than CD. It showed that the doping of Cu2+ could significantly reduce the compounding rate of photogenerated carriers by trapping electrons, resulting in an improved photocatalytic activity. It is worthy to note that the PL intensity of Ni-MOL is lower than that of CD(Cu)-Ni-MOL. This is because the energy band gap of Ni-MOL is 3.536 eV, which can only absorb ultraviolet light and does not produce electron transition under visible light irradiation, so PL has a low intensity. In addition, there are certain intrinsic relationships between the PL spectrum and photocatalytic activity of a semiconductor material according to the mechanisms of PL and photocatalysis. However, the inherent relationships are very complicated, and can properly be disclosed only on the basis of PL attributes, closely related to dopant species. For excitonic PL signals, the lower the PL intensity, the higher the separation rate of photo-induced charges and, possibly, the higher the photocatalytic activity (Zhang et al. 2000). The stronger the PL signal, the higher the content of surface oxygen vacancies and defects and possibly the higher the photocatalytic activity. Compared with Ni-MOL, the doped Cu in CDs and the loaded CD(Cu) in Ni-MOL could form more oxygen vacancies and defects. Therefore, the PL intensity of Ni-MOL was lower than that of CD(Cu)-Ni-MOL, while the photocatalytic performance of CD(Cu)-Ni-MOL was higher than that of Ni-MOL.

Photocatalytic mechanism

Figure S1 shows the effect of initial concentration on removal rate. As shown in Figure S1, it can be found that the value of C/C0 decreases with the increases of initial concentration. However, 83% removal capacity can still be maintained at the initial concentration 125 mg/L under visible light condition, indicating that CD(Cu)-Ni-MOL has excellent catalytic performance. Compared with visible light condition, the degradation performance of CD(Cu)-Ni-MOL decreased significantly under ultraviolet light conditions. It is based on the fact that the band gap of CD(Cu)-Ni-MOL is 2.071 eV (Figure 5(a) inset). 2.071 eV is corresponded to the absorption of visible light. Therefore, no catalytic degradation occurs under UV light. The existence of adsorption phenomenon leads to the decrease of C/C0 value with increases of initial concentration. In order to further investigate the adsorption progress, the adsorption equilibrium analysis was performed. The adsorption of TC onto CD(Cu)-Ni-MOL by different initial TC concentrations (25, 50, 75, 100, 125 and 150 mg/L) were shown in Figure S2. It is clear that the adsorption capacity of TC onto CD(Cu)-Ni-MOL was increased with the increase of initial TC concentrations. The remove rate at the initial TC concentration of 150 mg/L has reached 30.2%. The adsorption equilibrium was often described by adsorption isotherms. Freundlich isotherm models are the two most frequently-used models used to describe the adsorption isotherm process. Figure S3 shows the adsorption isotherms of TC with corresponding Langmuir and Freundlich plots. The fitted parameters of both Freundlich and Langmuir isotherms are shown in Table S1. The correlation coefficient of Freundlich model was higher than that of Langmuir model, indicating that isotherms results were better described by Freundlich model. In another word, the adsorption process was non-uniform and multilayer.

Pure Ni-MOL was not an efficient photocatalyst for TC degradation (Figure 6(a)). After loading CD on Ni-MOL, the removal rate of TC by CD-Ni-MOL reached 48.7%. The enhancement of photocatalytic performance was attributed to two facts. The first fact was that the broadband Ni-MOL could not absorb the visible light effectively. CD had the up-conversion luminescence property, which could effectively change the light absorption range of the composite photocatalyst. The second fact was that CD could act as both electron donor and acceptor, and act as carrier transferring bridge in the composite catalyst, of which could efficiently reduce the electron-hole pair compounding rate. It was interesting that the photocatalytic performance was further improved after loading Cu on the surface of CD-Ni-MOL. Based on FT-IR and XPS spectra, Cu had loaded on the surface of CD. The photo-generated electrons produced by CD could be rapidly transferred to Cu, avoiding the transfer to Ni located at Ni-MOL. Thus, it shortened the transfer distance. In addition, the electron transferred to Cu could also effectively avoid photoelectron-hole recombination. More importantly, Cu2+ may have been converted to Cu0 atom in photocatalytic reaction. To verify the existence of Cu0, a controlled experiment was designed. As depicted in Figure S4, the peak area ratio of Cu0 had been increased from 67.5% to 92.6% after illumination for 2 h, suggesting that the electrons produced by the photocatalytic process had reduced Cu2+ to Cu0. STEM-HAADF and EDS-mapping (Figure S5) further proved the existence of the Cu0 atom. Many studies have proved the existence of Cu0 could effectively improve the catalytic performance. In order to further investigate the effects of different metal ion species on catalytic performance, metal ions (Fe, Co, Ni, Cu and Zn) were loaded on CD. The results showed that the species of metal ions could affect the photocatalytic performance (Figure 6(b)). The effects of metals on photocatalytic performance improvement was Cu > Zn > Ni > Co > Fe. It was codetermined on the following factors. Firstly, different M-O or M-N bond (M: Cu, Zn, Ni, Co and Fe) possessed different electrical conductivity. Secondly, the doping of metal ions might induce the induction of electronic coupling in the newly formed intra-bandgap state proximity to the edge of the conduction band to induce electronic coupling of which prevented the charge complexation. Different species of metal ions equipped with different abilities of intra-bandgap state. Figure S6 shows the transient photocurrent response curves of CD(x)-Ni-MOL (X = Fe, Co, Ni, Cu, Zn). The photocurrent intensity followed the order of Cu > Zn > Ni > Co > Fe, indicating that the Cu induction of electronic coupling in the newly formed intra-bandgap state was near the edge of the Ni-MOL conduction band. Thirdly, different amounts of metal atoms were formed. The redox potentials of Cu2+/Cu, Zn2+/Zn, Ni2+/Ni, Co2+/Co and Fe2+/Fe were 0.340, −0.763, −0.236, −0.282 and −0.409 V, respectively. The higher the redox potential was, the easier it was to form metal atom by reduction reaction. Therefore, the photocatalytic performance reflected the following trend: Cu > Ni > Co > Fe > Zn. In addition, zinc-oxygen bond could generate photo-generated carriers after absorbing light, resulting in improved photocatalytic performance (Gotipamul et al. 2022). Finally, the photocatalytic sequence of different metal ions was in the order of Cu > Zn > Ni > Co > Fe.
Figure 6

Photocatalytic degradation dynamic curves of various materials, (a) Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL; (b) CD(x)-Ni-MOL, (X = Fe, Co, Ni, Cu and Zn).

Figure 6

Photocatalytic degradation dynamic curves of various materials, (a) Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL; (b) CD(x)-Ni-MOL, (X = Fe, Co, Ni, Cu and Zn).

Close modal

Figure S7 shows the effect of pH of degradation process. From Figure S7, it can be found that the removal rate reaches the maximum (62.4%) at pH = 8 under the dark condition, indicating that adsorption plays an important role in the degradation and removal of TC. Similarly, the maximum removal rate has reached to 98.3% at pH = 8 under the visible light. Compared with dark environment, the change of pH has less influence on removal rate under visible light. According to the adsorption-degradation theory, the initial stage of degradation removal is adsorption, and then the degradation stage. In the adsorption process, it is greatly affected by pH change, while in the degradation stage, it is fairly affected by pH change.

The trapping experiments were conducted to determine the predominant reactive species in photodegradation process (Figure 7). Herein, IPA, EDTA-2Na, K2Cr2O7 and p-BQ were selected to capture •OH, h+, e and , respectively. A significant decrease in the degradation rate of TC could be determined in the presence of p-BQ and EDTA-2Na relating to the addition of IPA and K2Cr2O7 to the reaction medium, indicating that the effective role of and h+ were predominant reactive species during the photodegradation TC progress.
Figure 7

Trapping experiments of the active species for the degradation of TC by CD(Cu)-Ni-MOL.

Figure 7

Trapping experiments of the active species for the degradation of TC by CD(Cu)-Ni-MOL.

Close modal

In order to further confirm the role of active free radical in the process of photocatalysis, electron spin resonance (ESR) technology was used to investigate the role of •OH and in the photocatalytic degradation of TC. As shown in Figure S8, the signals of DMPO- could be clearly observed. Compared with DMPO-, there was no signal for DMPO-•OH, suggesting that superoxide radicals indeed played a major role in the photocatalytic reaction. The results of ESR further verified the results of trapping experiments.

All samples showed sensitive photocurrent response in multiple on/off cycles as presented in Figure 8(a). Photocurrent intensity followed the order of CD(Cu)-Ni-MOL > CD-Ni-MOL > Ni-MOL. Ni-MOL exhibited a relatively low photocurrent density due to the accelerated electron-hole complexation. Generally speaking, it is believed that the stronger photocurrent signals imply more efficient transfer and separation of photoexcited electron-hole at the CD(Cu)-Ni-MOL interface. Meanwhile, the phenomenon that the photocurrent value remained stable after several cycles, indicating that CD(Cu) on Ni-MOL could effectively inhibit the electron hole recombination of the composite (Cen et al. 2021). The EIS spectra of Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL were used to study the transfer resistance of photoexcited holes from electrons (Figure 8(b)). Semicircles in Nyquist curves were used to characterize the process of charge transfer, while the diameters of semicircles represented the internal resistance of electrode-electrolyte in the electrolyte (Ma et al. 2014). It was observed that the arc radius on the EIS Nyquist plot of CD-Ni-MOL was smaller than that of Ni-MOL, suggesting that the CD-modified Ni-MOL had smaller electron transfer resistance to enhance the photocatalytic activity effectively. EIS performance showed that the arc radius of CD(Cu)-Ni-MOL was smaller than that of CD-Ni-MOL, confirming the successful sintering of Cu nanoparticles and CD precursors. It could be concluded that CD could be used as electron transfer mediators in vector electron transfer of Cu-CD-Ni-MOL, facilitating the interfacial charge separation and migration and effectively decreasing excited state burst and energy dissipation.
Figure 8

(a) Transient photocurrent response curves of Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL, (b) EIS of Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL.

Figure 8

(a) Transient photocurrent response curves of Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL, (b) EIS of Ni-MOL, CD-Ni-MOL and CD(Cu)-Ni-MOL.

Close modal

The bandgap energy of Ni-MOL was calculated as 3.536 eV as shown in Taucplot (Figure 6 inset). We plotted and analyzed the Mott-Schottky curve (Figure S9) with the purpose of identifying the semiconductor type and the flat charged potential of Ni-MOL. Ni-MOL was an n-type semiconductor and the flat-band potential was 0.063 eV (vs. Ag/AgCl). Therefore, the valence band potential of Ni-MOL could be identified as 3.257 V vs. NHE.

According to the above results, the reaction Equations (1)–(4) were speculated to elucidate the mechanism of TC degradation by CD(Cu)-Ni-MOL (Figure S10). Incidence of sufficient visible light to CD(Cu)-Ni-MOL caused the electron to excite from the valence band (VB) to the conduction band (CB) while the vacancy of electron was assigned to h+ Equation (1) (Ye et al. 2017). Then e was shifted to the surface of CD(Cu), while adsorbed O2 was reduced to radicals due to the change in the location of the valence band and conduction band of the composite caused by the addition of CD(Cu) Equation (2). Moreover, the excess electrons were transferred to Cu to reduce Cu2+ to Cu0 Equation (3). Then, the electrons transferred to Cu were reacted with dissolved oxygen to form Equation (4). At last, these reactive radicals ( and h+) completely mineralized the TC molecules. Consequently, not only can CD(Cu) dramatically promote the visible light utilization of CD(Cu)-Ni-MOL nanosheets through its unique upconversion properties, but it could also act as an acceptor for light energy generation electrons, effectively improving electron transfer in photocatalytic (Qin & Zeng 2017).
(1)
(2)
(3)
(4)

The stability of photocatalysts is a decisive factor in practical applications. Therefore, the stability of CD(Cu)-Ni-MOL was assessed, which showed the best degradation performance of TC with cyclic experiments in this study. Figure S11 shows that the decrease of the photo-catalytic activity of CD(Cu)-Ni-MOL in five cyclic experiments was tiny, suggesting that CD(Cu)-Ni-MOL could maintain a stable degradation activity in long-term use.

The HPLC-MS was applied to identify the intermediate products and to explore the degradation pathway (Figure S12). Interestingly, toroidal structures of tetracyclines were related to several readily ionizable functional and electron-rich groups (e.g. dimethylphenol groups and amino groups) that were susceptible to attack by reactive agents (Shen et al. 2022).

In pathway I, a double bond on P1 was oxidised by to produce a ketone group and a new hydroxyl group to give P2 (m/z = 477). Subsequently, compound P3 (m/z = 449) was obtained by demethylation as a result of attacking the position of the tertiary amine on the ring. Afterwards, It was possible to convert P3 to P4 (m/z = 393) by removing another methyl group and losing the amide group. It is mainly through open-loop and oxidation processes that the formation of P5 (m/z = 335) and P6 (m/z = 274) could be determined. The intermediate P6 was further degraded to small molecules (m/z = 102).

In pathway II, TC molecules were firstly converted to P7 (m/z = 427), which was ascribed to the dehydration products under the attack of h+. With the progression of the reaction, intermediates P8–P10 (with m/z = 301, 291, 207, and 149, respectively) were formed following terminal oxidative and ring-opening reactions triggered by •OH.

In this work, a novel 2D layered photocatalyst was successfully prepared by a one-step hydrothermal method. It exhibited excellent photocatalytic degradation ability of TC under visible light irradiation, with degradation rate reaching approximately 93.5% within 60 min. The results of UV-DRS, PL, transient photocurrent and EIS suggested that CD(Cu)-Ni-MOL could efficiently inhibit the recombination of photogenerated charge. In addition, CD could convert low-energy visible light to high-energy visible light through upconversion effect and the addition of Cu resulted in the change of carrier transport path and the formation of Cu0, which effectively improved the photocatalytic performance of Ni-MOL. h+ played a significant role and •O2 also played a role in the photo-degradation process. The TC degradation process and decomposition pathways were proposed based on HPLC-MS.

This work was financially supported by the Open Research Fund Program of Key Laboratory of Cleaner Production and Integrated Resource Utilization of China National Light Industry (CP2021YB10), Integration of Science, Education and Industry Innovation Pilot Project of Qilu University of Technology (Shandong Academy of Sciences) (2020KJC-ZD12), Cooperation Foundation for the Youth Scholars of Qilu University of Technology (Shandong Academy of Sciences) (2018BSHZ0022) and Natural Science Foundation of Shandong Province (ZR2017LEE024).

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Gotipamul
P. P.
,
Vattikondala
G.
,
Rajan
K. D.
,
Khanna
S.
,
Rathinam
M.
&
Chidambaram
S.
2022
Impact of piezoelectric effect on the heterogeneous visible photocatalysis of g-C3N4/Ag/ZnO tricomponent
.
Chemosphere
287
(
4
),
132298
.
Huang
Y.
,
Liang
Y.
,
Rao
Y.
,
Zhu
D.
,
Cao
J. J.
,
Shen
Z.
,
Ho
W.
&
Lee
S. C.
2017
Environment-friendly carbon quantum dots/ZnFe2O4 photocatalysts: characterization, bio-compatibility, and mechanisms for NO removal
.
Environ. Sci. Technol.
51
,
2924
2933
.
Johnson
J. A.
,
Zhang
X.
,
Reeson
T. C.
,
Chen
Y.-S.
&
Zhang
J.
2014
Facile control of the charge density and photocatalytic activity of an anionic indium porphyrin framework via in situ metalation
.
J. Am. Chem. Soc.
136
(
45
),
15881
15884
.
Li
S.
,
Shi
W.
,
Liu
W.
,
Li
H.
,
Zhang
W.
&
Hu
J.
2018
A duodecennial national synthesis of antibiotics in China's major rivers and seas
.
Sci. Total Environ.
615
,
906
917
.
Li
S. M.
,
Ji
K.
,
Zhang
M.
,
He
C. H.
,
Wang
J.
&
Li
Z. Q.
2020
Boosting the photocatalytic CO2 reduction of metal–organic frameworks by encapsulating carbon dots
.
Nanoscale
12
(
17
),
9533
9540
.
Ma
Y.
,
Xu
H.
,
Zeng
Y.
,
Ho
C. L.
,
Chui
C. H.
,
Zhao
Q.
,
Huang
W.
&
Wong
W. Y.
2014
A charged iridophosphor for time-resolved luminescent CO2 gas identification
.
J. Mater. Chem. C.
3
(
1
),
66
72
.
Pelaeza
M.
,
Nolan
N. T.
,
Pillai
S. C.
&
Seery
M. K.
2012
A review on the visible light active titanium dioxide photocatalysts for environmental applications
.
Appl. Catal. B Environ.
125
,
331
349
.
Wang
F.
,
Liu
Y.
,
Wei
H.
,
Wang
G.
,
Ren
F.
,
Liu
X.
,
Chen
M.
&
Volinsky
A. A.
2020a
Graphene induced growth of Sb2WO6 nanosheets for high-performance pseudocapacitive lithium-ion storage
.
J. Alloys Compd.
839
,
155614
.
Yang
W.
,
Wang
H. J.
,
Liu
R. R.
,
Wang
J. W.
,
Zhang
C.
,
Li
C.
,
Zhong
D. C.
&
Lu
T. B.
2021
Tailoring crystal facets of metal–organic layers to enhance photocatalytic activity for CO2 reduction
.
Angew. Chem. Int. Edit.
60
,
409
414
.
Zhang
W. F.
,
Zhang
M. S.
&
Yin
Z.
2000
Photoluminescence in anatase titanium dioxide nanocrystals
.
Appl. Phys. B
70
,
261
265
.
Zhu
Y.
,
Chen
J.
&
Kaskel
S.
2020
Porphyrin-based metal–organic frameworks for biomedical
.
Angew. Chem. Int. Ed.
60
(
10
),
5010
5035
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Supplementary data