Jiarong
Liang
a,
Wei
Li
ab,
Jianying
Chen
a,
Xiaoman
Huang
a,
Yingliang
Liu
ab,
Xuejie
Zhang
ab,
Wei
Shu
c,
Bingfu
Lei
*ab and
Haoran
Zhang
*a
aKey Laboratory for Biobased Materials and Energy of Ministry of Education, College of Materials and Energy, South China Agricultural University, Guangzhou 510642, P. R. China. E-mail: tleibf@scau.edu.cn; hrzhang@scau.edu.cn
bMaoming Branch, Guangdong Laboratory for Lingnan Modern Agriculture, Guangdong, Maoming 525100, P. R. China
cInstrumental Analysis & Research Centre, South China Agricultural University, Guangzhou 510642, P. R. China
First published on 26th October 2022
The covalent organic framework@carbon dot (COF@CD) composite was successfully constructed to achieve high-flux charge transfer and efficient photocatalytic activity for antibacterial photocatalytic therapy. Utilizing the establishment of intramolecular charge transfer between donor–acceptor (D–A) semiconductors and polymers with CO–H–N hydrogen-bonding groups, the charge transfer channel constructed by using a D–A COF semiconductor and hydrophilic CDs was built. This is the first report on hydrogen-bonded two-dimensional COF-zero-dimensional CDs for artificial antibacterial photosynthesis. In such a photocatalytic system, transient photovoltage (TPV) measurements demonstrated that CDs uniformly distributed on the surface of COF nanosheets play an important role in inhibiting charge recombination as both electron transfer and storage containers. Kinetic studies showed that the introduction of CDs greatly enhanced the charge separation efficiency by extracting abundant photogenerated π-electrons from the COF, resulting in the generation of more reactive oxygen species. COF@CDs (4 wt% CDs) present photocatalytic antibacterial activity with sterilization efficiency of over 95% in 1 h under visible light irradiation, with a decrease in the survival rate by 8.3 times compared to that of the COF. This result is attributed to the combined effect of the photoexcitation rate, carrier separation rate, and reduction rate of electrons accumulated in the CDs.
Compared with other photosensitizers, covalent organic frameworks (COFs), a novel class of crystalline porous organic semiconductors linked by light elements through strong covalent bonds, have attracted widespread attention in photocatalytic therapy, owing to their nontoxicity, broad absorption range, high photothermal stability, tunable chemical structure, and electronic properties.17,18 Consequently, building crystalline COFs with suitable photoactive components and layered superposition sequences is an exciting direction to realize a new generation of highly active artificial photosystems. This ordered structure facilitates customized and regulated electronic structures and chemical functions of COFs at the molecular level. The strong photoelectric conversion ability of COFs makes them a promising material in the field of photocatalysis. However, there is still an inherent problem in COFs, the exciton binding energy of COFs is relatively high compared to that of inorganic semiconductors.19 This means that the separation and migration rates of photogenerated electron–hole pairs in pure COFs are limited only by changing the structure and tuning the bandgap.20,21 Therefore, the maximum extraction of carriers from photoexcited COFs is the most important concern in efficient photoredox catalysis. In contrast to the traditional methods for forming heterojunctions with metals and semiconductors, Long et al. engineered a metal–insulator–semiconductor (MIS) photosystem based on donor–acceptor (D–A) type COF photocatalysts with tunable photoelectronic conduction.22 In particular, with the inherent different electron affinities under light irradiation, such a D–A type COF link by an imine conjugated polymer skeleton, has intramolecular charge transfer (ICT) from D to A, and the electrons accumulate preferentially in receptors with efficient electron mobility through the delocalization of π-electrons.23 To achieve more efficient and faster carrier transfer, it is necessary to induce an internal driving force on the two-dimensional semiconductor plane to delocalize the optical carriers.
Based on the above, we constructed a binary system with a D–A type COF and selected carbon dots (CDs) as both the electron extractant and storage media. CDs are ideal candidate conductive media, owing to their high-water solubility and outstanding electron transfer and sinking efficiency, which can improve the dispersibility and photogenerated carrier separation efficiency of COFs.24,25 Considering the structural characteristics of COF semiconductors, the uniform distribution of CDs, with strong chemical affinity on the surface of COFs (C, N, and O), is conducive for π-electron emission from the skeleton of COFs and maximizing the π-electron traversing efficiency.
In this study, an intramolecular charge transfer system was constructed to extract π-electrons from photosensitive D–A COF semiconductors to CDs through hydrogen bonding, and a nanostructured photosystem was designed to demonstrate a high-efficiency bactericidal mechanism. The incorporation of CDs into a D–A type COF to construct antibacterial photocatalyst composites has not been reported to date. As illustrated in Scheme 1, under photoexcitation, photogenic electrons are circulated and transmitted from 4,4′,4′′-(1,3,5-triazine-2,4,6-triyl) trianiline (TTA), an electron donor, to 1,3,5-triformylphloroglucinol (Tp), an electron acceptor. After Tp accepts the electrons, the remaining electrons flow back to the donor and rapidly recombine with the photogenic hole, with only a few electrons escaping. The amount of ROS produced is restricted. In contrast, with the loading of CDs, a charge transfer channel made of = CO–H–N = hydrogen bonds is formed between the COF and CDs. The photogenerated electrons trapped within the porous structure of the COFs are extracted by CDs, causing electrons to accumulate in the CDs for more ROS evolution. The generated electrons (e−) reduce dissolved oxygen (O2) to a superoxide radical (˙O2−) and hydrogen peroxide (H2O2), the electrons further reduce H2O2 to hydroxyl radicals (˙OH), and ˙O2− can further be oxidized by holes (h+) to 1O2. Our study indicated that the prepared COF@CDs as APCT had better catalytic efficiency against Escherichia coli (E. coli) than a pure COF, and the in vitro disinfection efficiency was >95% upon activation with visible light for 60 min.
Scheme 1 Schematic illustration of a photocatalytic antibacterial system over the COF and COF@CD composite. |
To gain insight into the surface morphology and microstructure, scanning electron microscopy (SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM) were exploited. The SEM images of the COF and COF@CDs were shown in Fig. S3(a and b)†, the overall morphology shows a two-dimensional sheet-like COF stacked layer by layer. The COF exhibited microstructures with ultrathin 2D layered π-interaction structures, as shown in Fig. 2a. High resolution electron transmission microscopy imaging (HR-TEM) (Fig. 2b), local magnified imaging (Fig. 2d) and fast Fourier transform (FFT) images (Fig. 2c) show the highly ordered and periodic structure of the COF with a lattice spacing of 0.24 nm, corresponding to the excellent crystallinity of XRD. As shown in Fig. 2f, the CDs have a regular spherical structure, with uniform dispersion without aggregation and a lattice spacing of 0.19 nm as clearly shown in Fig. 2h. Significantly, from the TEM imaging of COF@CDs in Fig. 2i, the individual crystallites highlighted by dot circles are uniformly distributed on the surface of the COF nanosheet, displayed an interplanar spacing of 0.19 nm as shown in Fig. 2g, which is ascribed to CDs. This intimate interface facilitates charge conduction for carrier separation in photocatalysis. The sizes of the COF and CDs were measured by AFM (Fig. 2e–j). The COF exhibited an ultrathin nanosheet structure and its average height and width were approximately 1.5 and 100 nm. The CDs were uniformly spherical with an average height of 2 nm. These morphological results prove that 0D CDs can be easily deposited onto a two dimensional COF nanosheet through strong hydrogen bonds.
Additionally, antibacterial activity was investigated using the fluorescent live/dead staining method with confocal laser scanning microscopy (CLSM). Live and dead bacteria were stained with fluorescein diacetate (FDA) for green fluorescence and propidium iodide (PI) for red fluorescence. As shown in Fig. 3e, both E. coli cultures, except for the COF and COF@CDs treated with light groups showed bright green emission and no red fluorescence, indicating that the bacteria remained unaffected by the experimental treatments. As expected, a large amount of green fluorescence and scattered red fluorescence suggested partial membrane damage in the COF-treated bacteria under light exposure. The emission of red fluorescence increased significantly, and no green fluorescence was observed, indicating that the number of dead cells increased dramatically, demonstrating the severe membrane disruption capacity of COF@CDs under visible light. This result further highlights the synergistically enhanced antibacterial effect of the CD-integrated COF. The change in surface morphology of bacteria, as shown in the SEM images (Fig. 3f), reflects the influence of the samples on bacteria at the microscopic level. The smooth and intact cell walls of rod-shaped E. coli were damaged and roughened seriously by high concentrations of the COF and COF@CDs under light treatment. The cell structure of E. coli in the control group with light irradiation and the treatment group in the dark was intact without obvious shrinkage or damage, consistent with the results of the fluorescent live/dead staining method. Therefore, both the COF and COF@CD nanocomposites have antibacterial effects under light, but the composites have a higher sterilization efficiency than the pure COF. Photocatalytic antibacterial stability for COF@CDs was evaluated and is shown in Fig. S4.† There was no obvious decay of antibacterial activity of COF@CDs for two cycles under light irradiation. After irradiation and continuous agitation for 120 min, the antibacterial activity of COF@CDs began to decrease and the cell viability at 180 min was close to 40%. However, although the antibacterial efficiency for the fourth cycle was not as good as that for the first cycle, the sterilization rate maintains a similar level as the third cycle at 240 min, demonstrating the good photocatalytic antibacterial stability of COF@CDs.47,48
We further used electron paramagnetic resonance spectroscopy (EPR) to detect the types and intensity of ROS with 5,5-dimethyl-1-pyrroline N-oxide (DMPO) and 2,2,6, 6-tetramethyl-1-piperidine (TEMP) as spin probes. As shown in Fig. 4b, regardless of the type of ROSs (DMPO-˙OH, DMPO-˙O2−, and TEMP-1O2), both the COF and COF@CDs could generate ROS under illumination, and the characteristic peak intensity of COF@CDs was higher than that of the COF. There were no signals of ROS to be detected for samples in the dark, revealing that both ˙OH, ˙O2−, and 1O2 were excited by visible light. This measurement further confirmed that both the COF and COF@CDs have a high ROS production efficiency and that COF@CDs have a better production efficiency.
To explore which reactive oxygen species plays a major role in the photocatalytic disinfection of COF@CDs, ROS quenching experiments were performed. ˙OH, ˙O2−, and 1O2 were quenched with isopropanol (IPA), TEMPO and L-histidine (L-his), respectively. As shown in Fig. 4c, compared to the control group, the cell viability significantly increased by 41–57% for COF@CDs with the introduction of ROS scavengers, indicating that ROS were the dominant antibacterial factors and multiple reactive oxygen species cooperated in the photocatalytic disinfection system. The inset intuitively displays the E. coli colony growth. We then used p-nitrophenol, an electron scavenger, and p-methoxyphenol, a hole scavenger, in ROS-quenching experiments, to explore the role of electrons and holes generated by COF@CDs in ROS production. As shown in Fig. 4d, the fluorescence intensity of DCF was greatly reduced in the presence of p-nitrophenol, while there was a weak enhancement in the presence of p-methoxyphenol. This result reveals that electrons play a leading role in ROS production.17
The UV-vis diffuse reflectance spectrum (DRS) measurements in Fig. 5a show the light absorption properties of the as-prepared samples. The COF exhibits broad optical absorbance, even covering the entire UV-vis region, resulting from the great enhancement of the delocalization of π-electrons within the 2D skeleton by the highly planar π-conjugated structure. CDs have a wider absorbance range, with strong absorbance extending into the near-infrared region. Compared to the pure COF, the absorption edge of the COF@CD composite did not shift noticeably, suggesting that the CDs were not incorporated into the COF lattice but were loaded on the COF surface, which is consistent with the XRD results. The corresponding optical band gap of the COF was calculated using the Kubelka–Munk equation: α(hν) = A(hν − Eg)1/2 (where α, ν, h, A, and Eg represent the absorption coefficient, light frequency, Planck's constant, constant value, and band gap energy, respectively).32 As shown in the inset of Fig. 5a, the optical band gap of the COF calculated from the Tauc plot was approximately 2.18 eV. According to the Mott–Schottky plots (Fig. 5b) measured at the frequencies of 1000, 2000, and 3000 Hz, the positive slope of the linear plots indicate that the TTATp-COF is an n-type semiconductor, and the derived flat band (FB) potential of the COF is approximately −0.89 eV vs. Ag/AgCl.33 For an n-type semiconductor, the bottom of the conduction band (CB) potential is generally 0.2 eV, which is more negative than the FB potential, and hence, the CB potential of COF nanosheets is −1.09 eV vs. Ag/AgCl, that is −0.89 eV vs. NHE, according to the equation, ENHE = EAg/AgCl + EAg/AgClθ; EAg/AgClθ = 0.1976 eV. Considering the Eg value of 2.18 eV, estimated from the UV-vis DRS, the valence band (VB) should be 1.38 eV. By calculating the results of VB-XPS in Fig. 5c, it can be directly obtained that the VB position of the TTATP-COF is located at 1.34 eV, which is almost the same as the above results.
To better understand the improved interface charge transfer efficiency, the band alignment of the COF and CDs was analyzed using ultraviolet photoelectron spectroscopy (UPS) measurements. As shown in Fig. 5d and e, the UPS results show that the secondary electron cutoff values of the COF and CDs are 17.9 and 20.10 eV; by subtracting the He I excitation energy (21.2 eV), the work functions are calculated to be 3.3 and 1.1 eV, respectively. Because of the smaller work function of the COF, the photogenerated electrons on the COF nanosheet are spontaneously transferred to CDs, driven by thermodynamics. An electron transfer pathway from the COF to CDs was built.34
Based on the above measurement results, we calculated the band distribution alignment of COF@CDs and proposed a possible production process of reactive oxygen species in APCT, as shown in Fig. 5f. When excited with photon energies exceeding the bandgap, the photogenerated electrons on the surface of the COF make transitions and are captured by CDs, leading to effective separation of electron–hole pairs. Since the CB potential of the COF (−0.8 eV vs. NHE) was more negative than the redox potential of O2/˙O2− (−0.33 eV vs. NHE), the electrons (e−) of the CB could reduce O2 adsorbed on the surface of COF@CDs to ˙O2−, which could be further oxidized by holes (h+) leaving the VB of the COF to 1O2.35 Furthermore, the COF has sufficient reduction ability to reduce O2 to H2O2 under irradiation (O2/H2O2 = +0.28 eV vs. NHE).36,37 In addition, the VB potential of the COF (+1.38 eV vs. NHE) was positive compared to that of the oxidation potential of OH−/˙OH (+1.99 eV vs. NHE), and it was thermodynamically forbidden for h+ to oxidize H2O to ˙OH.38 However, ROS detection measurements have demonstrated that the COF and COF@CDs can produce ˙OH under visible light irradiation, which has been attributed to the decomposition of H2O2 into ˙OH (+0.73 eV vs. NHE). The photoinduced electrons on the CB of the COF are transferred to CDs, which serves as an electron conduction and storage medium, efficiently decelerating the recombination of electron–hole pairs and prolonging the lifetime of charge carriers, resulting in increased ROS production, as shown in the following formulae eqn (1)–(5).
COF@CDs + hv → e− + h+ | (1) |
O2 + e− → ˙O2− | (2) |
˙O2− + h+ → 1O2 | (3) |
˙O2− + e− + 2H+ → H2O2 | (4) |
H2O2 + e− → ˙OH + OH− | (5) |
The charge carrier dynamics are also reflected by steady-state photoluminescence (PL) spectra and time-correlated single-photon counting (TCSPC) measurements. As shown in Fig. 6a, the COF exhibited characteristic emission peaks at 623 nm. When CDs were immobilized on the COF, significant quenching of the emission was observed. There is a charge transfer channel between the COF and CDs, which can facilitate photoinduced electron transfer from the excited state of the COF to CDs, and the charge density is redistributed, leading to effective suppression of the photoexcited carrier recombination of the COF. We further analyzed the inactivation of the stimulated COF by monitoring the emission decay. As shown in Fig. 6b, COF@CDs exhibited longer excited-state lifetimes (2.57 ns) than the COF (2.34 ns). CDs can efficiently accept electrons from the COF and retard carrier recombination.39 It is worth noting that the energetic π-electrons trapped in the COF skeleton can smoothly traverse to CDs, consequently achieving more efficient electron–hole pair migration and a larger photoexcitation rate. To further confirm that the CDs had excellent electron transfer capability, transient photocurrent response and electrochemical impedance spectroscopy (EIS) of different samples were investigated. As shown in Fig. 6c, the strength of the transient photocurrent response intensity for several on–off cycles of COF@CDs was much higher than that of the pure COF, which was beneficial to the redox reaction for ROS production.40 Similarly, from the EIS in Fig. 6d, higher electronic conductivity can be seen based on the smaller Nyquist plot radius for COF@CDs than for the COF during charge transfer processes, implying a decrease in charge transfer resistance.41 In Fig. S5,† the polarization curve shows the maximum overpotential of the pure COF. After loading CDs, the overpotential decreased, indicating that the CDs can accelerate the separation of photogenerated carriers.
To study the dynamic information for delivery kinetics of photoinduced carriers, transient photovoltage (TPV) measurements were performed. As an important detection technique, TPV can directly reflect the photophysical process under working conditions, including the generation, separation, transport, accumulation, and recombination of photoinduced internal carriers. Fig. 7a shows the original TPV curves of the COF and COF@CDs. The transient photovoltage curve was positive for COF@CDs, but negative before 0.079 ms, and then became positive for the COF. A negative photovoltage response indicates electron accumulation on the surface, and a positive value suggests holes accumulation on the surface. It means that with irradiation, the electron–hole pairs were separated and then electrons escaped to the surface from the inside of the COF, showing a negative voltage. After 0.079 ms, the electrons were consumed by the reaction with oxygen, leaving the photogenerated holes that gradually accumulated at the interfaces of the COF, showing a positive voltage. The obvious difference for COF@CDs was that after loading CDs, a large number of electrons were rapidly extracted and stored by CDs, while the holes accumulated on the COF nanosheets, resulting in a positive signal on the TPV curve.42 As presented in Fig. 7b, the maximum photocarrier extraction efficiency is expressed by the shaded area. The value of COF@CDs is much larger than that of the COF, revealing that the CDs have an excellent charge storage capacity. The photocatalytic performance of charge extraction by CDs is superior to that of the COF.43–45 Here, we define an equation Tt/Tr = L to clarify the measurement of TPV,46 where, Tt is the time at which the photocarrier concentration reaches its maximum, Tr is the time required for complete recombination of the photoinduced carriers, and L is the ratio of 0 to 1. The smaller the L value, the faster the transfer and the longer the lifetime of the photogenerated carriers. The amplification diagram of Tt is shown in Fig. 7c, in which the Tt of COF@CDs is faster than that of the COF, illustrating the faster charge extraction rate of COF@CDs. The values of Tt, Tr and L for the COF and COF@CDs are listed in Table 1. The results indicated that COF@CDs had a smaller L value (0.0028) than the COF (0.004), suggesting a faster separation rate of electron–hole pairs and longer lifetime of charge carriers. CDs play an important role in efficiently extracting and storing electrons to accelerate electron transfer and prevent electron–hole recombination, thereby improving the process of APCT.
Fig. 7 (a) The transient photovoltage curves, the corresponding (b) amount of charge extraction and (c) maximum charge extraction time of the COF and COF@CDs. |
Samples | T t (ms) | T r (ms) | L |
---|---|---|---|
COF | 0.0022 | 0.5495 | 0.004 |
COF@CDs | 0.0019 | 0.6761 | 0.0028 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ta03978h |
This journal is © The Royal Society of Chemistry 2022 |