Iqra Rabania,
Nguyen Tien Tranbc,
Muhammad Faheem Maqsoodd,
Mosab Kaseeme,
Ghulam Dastgeer
f and
Hai Bang Truong
*gh
aAntwerp Engineering, Photoelectrochemistry and Sensing (A-PECS), University of Antwerp, Groenenborgerlaan 171, 2020 Antwerp, Belgium
bCenter for Advanced Chemistry, Institute of Research and Development, Duy Tan University, 03 Quang Trung, Da Nang 550000, Vietnam
cFaculty of Natural Sciences, Duy Tan University, 03 Quang Trung, Da Nang 550000, Vietnam
dSchool of Engineering, The Australian National University, Canberra, Australian Capital Territory 2601, Australia
eDepartment of Nanotechnology and Advanced Materials Engineering, Sejong University, Seoul 05006, Republic of Korea
fDepartment of Physics and Astronomy, Sejong University, Seoul 05006, South Korea
gOptical Materials Research Group, Science and Technology Advanced Institute, Van Lang University, Ho Chi Minh City, Vietnam
hFaculty of Applied Technology, Van Lang School of Technology, Van Lang University, Ho Chi Minh City, Vietnam. E-mail: truonghaibang@vlu.edu.vn
First published on 26th August 2025
The removal of pharmaceutical and organic contaminants from wastewater remains a pressing challenge for conventional treatment technologies. In this study, a novel photocatalyst composed of Zeolitic Imidazolate Framework-8 (ZIF-8) nanocrystals integrated with a graphene oxide (GO) matrix was developed via a facile interfacial synthesis approach. The structural and morphological properties of the resulting ZIF-8/GO composite were characterized using X-ray diffraction, X-ray photoelectron spectroscopy, and transmission electron microscopy. The composite's photocatalytic efficiency was assessed through the degradation of Rhodamine B (Rho-B), a representative dye pollutant, and 5-fluorouracil (5-Flu), a widely used pharmaceutical, under visible light irradiation. The ZIF-8/GO catalyst achieved near-complete degradation of Rho-B (100%) and 5-Flu (97.4%) within 100 minutes. This high performance is attributed to the synergistic effects of ZIF-8 and GO, which enhance surface area, improve light absorption, and promote charge separation. Moreover, the catalyst retained considerable activity over five cycles, with only a 12.8% decrease in Rho-B degradation due to minor material loss. The findings demonstrate that the ZIF-8/GO composite is a highly efficient, recyclable, and sustainable photocatalyst, showing great promise for the removal of complex pollutants from wastewater and contributing to environmentally friendly water purification technologies.
In addition to dye pollution, pharmaceutical pollutants—particularly anticancer drugs—have become emerging contaminants of global concern.9,10 Studies indicate that the worldwide consumption of anticancer drugs exceeds 75 tons per year, with a considerable fraction excreted unmetabolized into sewage systems.11–15 Significant research efforts have shown that anticancer pharmaceuticals, even at trace concentrations in the ng L−1 to μg L−1 range, exert cytotoxic effects on aquatic organisms, including impaired reproduction, growth inhibition, and genotoxicity.16,17 Although these drugs are formulated to disrupt RNA or DNA synthesis in malignant cells, their stability and toxicity render them hazardous to aquatic ecosystems.18 Consequently, there is an urgent need to develop efficient strategies for eliminating anticancer agents from water bodies. These compounds, aimed at inhibiting cellular proliferation, are typically resistant to biodegradation and poorly removed by conventional biological treatment processes.19
Advanced oxidation processes (AOPs) have emerged as a viable alternative for degrading such contaminants, leveraging reactive species to break down pollutants.20–22 A wide range of AOPs, including ozonation, UV/chlorine, photo-Fenton, sonolysis, wet oxidation, and photocatalysis, as well as integrated hybrid approaches, have been employed to degrade organic contaminants.23–28 Among these, photocatalysis is predictable as eco-friendly remediation methods, garnering significant attention.1,29–31 This technique has been employed to degrade the organic pollutants from wastewater by producing highly reactive chemical species that can mineralize a wide range of organic contaminants.32 However, conventional photocatalysts often exhibit drawbacks such as broad bandgaps and limited surface areas, which hinder their practical efficiency in water treatment applications.8,30,33 These drawbacks necessitate the development of advanced photocatalytic materials that can overcome these challenges and provide enhanced performance in the degradation of both pharmaceutical and organic pollutants.
In this context, metal–organic frameworks (MOFs) have emerged as a class of crystalline porous materials composed of metal clusters coordinated with organic ligands.4,34 By carefully selecting organic and inorganic components, MOFs can be engineered to have specific properties and architectures.35 In recent years, MOFs have emerged as versatile materials with different implementations, including drug delivery gas capture and separation technologies, as well as light-driven catalytic processes.36–38 A subclass of MOFs, Zeolitic Imidazole Frameworks (ZIFs), such as ZIF-8, consists of organic linkers and coordinated inorganic clusters. ZIF-8 is particularly notable for its ability to act as both a sacrificial template and a metal precursor.39,40 Its attributes, such as a high surface area, adjustable porosity, and superior chemical durability, (with pore sizes around 1.16 nm and a pore volume of 0.60 cm3 g−1), make ZIF-8 especially attractive for applications in photocatalysis.41–43
Specifically, ZIF-8 has been employed for drug delivery in the treatment of cancer and bacterial infections. However, ZIF-8 has limitations, including poor light absorption, rapid recombination of photogenerated electron–hole pairs, and excessive release of Zn2+, which can lead to reduced photocatalytic performance and cytotoxicity.44 Forming composites with other materials has been a promising solution to these issues. In this regard, certain two-dimensional materials like GO,45 Ti3C2Tx46 and MoS2 nanosheets47 have demonstrated excellent photocatalytic properties. These materials can form heterojunctions with semiconductors, effectively tuning their bandgaps and thereby enhancing their photocatalytic performance.
Based on the versatile features of GO and ZIF-8, they are ideal candidates to combine using a cost-effective solvothermal strategy, making them effective photocatalysts. The synergistic effect between GO and ZIF-8 contributes to the formation of a well-defined, uniform composite structure with enhanced surface area, leading to improved charge separation. This study aims to: (1) synthesize a homogeneous ZIF-8 layer on a GO substrate to attain high surface area and porosity, (2) assess the photocatalytic performance of ZIF-8@GO utilizing dye (Rho-B) and drug (5-Flu) as model contaminants, (3) examine the influence of operational parameters on photocatalytic activity, (4) test the reusability and stability of the composite over five consecutive cycles, and (5) elucidate the degradation pathways of Rho-B and 5-Flu.
To assess the durability and reuse potential of the catalyst, the used material was recovered, thoroughly rinsed with deionized water about five times, and reused in subsequent experiments. Additionally, to elucidate the role of reactive oxygen species (ROS) in the degradation mechanism, trapping experiments were performed. These tests involved adding specific scavengers to the reaction mixture to quench different ROS. Furfuryl alcohol (FFA) was used to quench hydroxyl radicals (˙OH), p-benzoquinone (BQ) was used to quench superoxide radicals (˙O2−), and ethylenediaminetetraacetic acid disodium salt (EDTA-2Na) was used to quench photogenerated holes (h+). By introducing these scavengers and monitoring the changes in the degradation rate, the contribution of each type of reactive species to the photocatalytic process could be determined. This approach helps to understand which ROS are most active in the degradation mechanism.
To further assess the crystal distribution within the ZIF-8/GO composite catalyst, elemental mapping was conducted, with the results shown in Fig. 1(d–h). The analysis confirms a homogeneous distribution of elements throughout the composite, indicating a successful integration of the materials. This uniform dispersion contributes to enhanced functional properties. The interaction between the GO matrix and ZIF-8 particles results in a composite with superior performance characteristics, making it well-suited for high-value photocatalytic applications. Additionally, the EDS spectrum in Fig. 1(i) shows elemental peaks for carbon (C) and oxygen (O) from GO, along with zinc (Zn) and nitrogen (N) from ZIF-8, confirming the successful synthesis and integration of GO and ZIF-8 in the composite.
To further examine the crystal structure, XRD analysis of the GO, ZIF-8, and the ZIF-8/GO composite was performed, with the resulting patterns shown in Fig. 1(j). Pristine GO (Fig. S2) displayed a prominent peak at a 2θ angle of 11.1°, which corresponds to the (001) plane.48 This peak reflects the interlayer spacing between GO sheets and confirms the presence of oxygen-containing functional groups. Fig. S3 presents the XRD pattern for pristine ZIF-8 nanocrystals. ZIF-8 typically exhibits diffraction peaks at 7.02°, 10.2°, 12.6°, 14.6°, 16.4°, and 17.9°, which are consistent with the reported crystallographic data for this framework.8,45 The XRD pattern of the ZIF-8/GO composite (Fig. 1(j)) shows diffraction peaks at 7.02°, 9.7°, 11.1°, 13.0°, 14.1°, 16.4°, and 17.9°, confirming the preservation of the crystalline structure of ZIF-8 within the composite. Importantly, the GO peak at 11.1° remains visible in the composite, although with a slight change in sharpness compared to pristine GO. This observation suggests that while ZIF-8 nanocrystals are successfully integrated with the GO sheets, the layered structure of GO is still partially retained. Minor shifts in some ZIF-8 peaks further indicate interactions between the two components, leading to subtle distortions in the crystal lattice.
To further investigate the surface features, BET measurements were carried out to measure the photocatalysts' specific surface area, with the results illustrated in Fig. 2(a and b). The N2 adsorption/desorption isotherms of all catalysts show similar hysteresis loops, indicating comparable porosity, as depicted in Fig. 2(a and b). According to IUPAC classification, all catalysts exhibit type-I isotherms, suggesting strong interactions between the adsorbent and adsorbate.49,50 However, GO demonstrates a type-IV isotherm with an H3-type hysteresis loop under the same conditions,51 as shown in Fig. 2(a). Based on the calculations of the BET measurements, the ZIF-8/GO composite showed highest surface area (2230 m2 g−1), exceeding those of GO (720 m2 g−1) and ZIF-8 (1190 m2 g−1). In addition, BJH calculations revealed pore sizes of 1.134 nm (GO), 1.0134 nm (ZIF-8), and 0.6434 nm (ZIF-8/GO). Corresponding pore volumes were calculated as 0.026, 0.204, and 0.328 cm3 g−1. The superior photocatalytic efficiency of the ZIF-8/GO composite is due to its larger surface area and mesoporsity, which promote efficient electron–hole pair separation. Thus, ZIF-8/GO composite was chosen as the optimized photocatalyst for further studies due to its high surface-to-volume ratio.
![]() | ||
Fig. 2 BET analysis including (a and b) isothermal and BJH profiles for GO, ZIF-8 and ZIF-8/GO composite. |
High-resolution XPS analysis was conducted to confirm the chemical composition and valence states of GO, ZIF-8, and the ZIF-8/GO composite, with results shown in Fig. 3. The ZIF-8/GO composite primarily consists of C, O, N, and Zn elements. In the C 1s spectrum of GO (Fig. 3(a)), peaks corresponding to various carbon–oxygen bonds were identified: C–C (sp2 carbon) at 284.5 eV, O–CO at 289 eV and C–O at 286.5 eV. ZIF-8 (Fig. 3(b)) showed peaks at 284.5 eV for C–C, 285.6 eV for C–N, and 287.8 eV for C
O.34,52 The GO@ZIF-8 composite (Fig. 3(c)) exhibited peaks from both GO and ZIF-8, with slight shifts in XPS peaks attributed to synthesis conditions and crystal structure distortions.
The N 1s spectra for GO, ZIF-8, and the ZIF-8/GO composite reveal nitrogen-containing functional groups, with deconvolution showing Pyridinic N, Pyrrolic N, and Graphitic N peaks, corresponding to different chemical environments4,53 (Fig. 3(d–f)). These nitrogen species influence the material's electronic structure and energy storage potential. The high-resolution Zn 2p XPS profiles for ZIF-8 and ZIF-8/GO (Fig. 3(g and h)) display Zn 2p3/2 at 1045.1 eV and Zn 2p1/2 at 1021.8 eV, indicating Zn2+ ions within the ZIF-8 framework, verifying the effective integration of ZIF-8 into the GO matrix.54 The peak separation observed in the Zn 2p XPS spectra was 22.9 eV for ZIF-8 and slightly increased to 23.09 eV for the ZIF-8/GO composite. This small shift in peak separation suggests subtle changes in the electronic environment of the Zn2+ ions, likely due to the interaction between ZIF-8 and the GO matrix during composite formation.
Initially, the photocatalytic performance of GO and ZIF-8 alone was assessed to establish a baseline for comparison. To assess the effect of light alone, the degradation of Rho-B was also monitored under visible light for 30 minutes without adding any photocatalyst. As shown in Fig. 4(a), the absorbance spectrum exhibited negligible change, indicating that Rho-B remained stable and did not undergo photodegradation in the absence of a photocatalyst.
To further investigate the adsorption–desorption equilibrium, experiments were conducted in the dark for 30 min using GO, ZIF-8 and ZIF-8/GO catalysts. The results, depicted in Fig. S4–S6, show that the absorbance intensity of Rho- B remained stable with GO and ZIF-8/GO catalysts, indicating no adsorption or desorption of the dye on these materials. In contrast, slight degradation of the absorbance spectra was observed with the ZIF-8, suggesting some level of interaction with the dye.
The absorbance intensity of Rho-B with GO showed a similar pattern when exposed to visible light for the 40 min, followed by a gradual decline over the next 100 min at the given interval, as illustrated in Fig. 4(b). This indicates that GO facilitates the photocatalytic degradation of Rho-B, but slowly. In the case of ZIF-8, a similar behavior was observed where the absorbance intensity of Rho-B decreased steadily under visible light for 100 min, as shown in Fig. 4(c). This suggests that ZIF-8 also contributes to the photocatalytic degradation of Rho-B, though at a consistent rate over time.
Upon introducing the ZIF-8/GO photocatalyst, a pronounced and continuous decline in the absorbance intensity of Rho-B was recorded from the onset of visible light irradiation up to 100 min, as seen in Fig. 4(d). This significant and sustained decrease indicates that the ZIF-8/GO catalyst possess enhanced photocatalytic activity compared to the individual GO and ZIF-8 catalysts. This superior performance is likely due to the synergistic effects of the core–shell configuration, which improves the efficiency of the photocatalysis.
To quantify this performance, the degradation efficiency of the photocatalysts was calculated using the formula: η(%) = (Co − Ct)/Co × 100,57 where C represents the initial concentration (at time t = 0), and Ct denotes the concentration at a given time t under visible light. Among the tested materials, the ZIF-8/GO composite demonstrated a notably high degradation efficiency for Rho-B, achieving 100% degradation within 100 min. This performance is significantly higher compared to GO (37.4%) and ZIF-8 (51.8%), as shown in Fig. 4(e).
Additionally, a linear kinetic relationship was assessed by fitting the degradation data to the pseudo-first-order model using the equation: ln(Co/Ct) = −Kappt.58,59 Here, Co represents the initial concentration of Rho-B, and Ct is the concentration at a specific time t. The linear fits of the data exhibited correlation coefficients (R2) greater than 0.98, indicating a strong linear relationship. This suggests that the degradation of Rho-B adheres to pseudo-first-order reaction kinetics. The kinetic constant (kapp) for the degradation of Rho-B using the ZIF-8/GO catalyst was determined to be 0.0258 min−1. This value is significantly higher than those for GO (0.0048 min−1) and ZIF-8 (0.0074 min−1), as depicted in Fig. 4(f). These findings underscore the superior catalytic efficiency of the ZIF-8/GO catalyst compared to the other tested materials.
The enhanced photocatalytic activity of the ZIF-8/GO catalyst can be attributed to several synergistic mechanisms: (1) the combination of ZIF-8 and GO results in an increased surface area and more active sites for the adsorption and degradation of pollutants like Rho-B, (2) ZIF-8 generates electron–hole pairs (e−/h+) under visible light, but these pairs tend to recombine, reducing efficiency and GO, acting as an excellent electron acceptor and transporter, captures and transports the photo-generated electrons from ZIF-8, thereby reducing recombination rates. This efficient charge separation and transfer mechanism allows more holes to remain available in ZIF-8 for the oxidation of pollutants, significantly enhancing the photocatalytic activity. Additionally, (3) the combination of ZIF-8 and GO broadens the light absorption spectrum, increasing the number of generated charge carriers. The configuration of ZIF-8/GO maximizes the contact area and ensures efficient electron transfer between the two materials, leading to improved photocatalytic performance.
In practical applications, the recyclability of a catalyst is a crucial indicator of its stability.2 Therefore, the ZIF-8/GO catalyst was subjected to five successive photocatalytic cycles under consistent experimental conditions to evaluate its reusability. In each round, the optimized ZIF-8/GO material was applied to the degradation of Rho-B under visible light. Following each cycle, a 2 mL aliquot of the reaction mixture was taken, and the catalyst was separated from the Rho-B solution by centrifugation. The recovered catalyst was thoroughly rinsed with deionized water several times and then reused in the subsequent cycle under identical photocatalytic conditions. During the initial cycle, the ZIF-8/GO composite achieved complete degradation (100%) of Rho-B within 100 minutes of visible light exposure. For the subsequent cycles (second, third, fourth, and fifth), the degradation efficiencies were 95.7%, 90%, 89.3%, and 87.0%, respectively, as depicted in Fig. 4(g). Notably, even after five cycles, the degradation efficiency remained high at 87.0%, indicating that the ZIF-8/GO catalyst maintains substantial photocatalytic activity. The observed decrease in catalytic performance of approximately 13% in the fifth cycle is attributed to the loss of catalyst during the washing process.
The loss of photocatalyst during the experiments was calculated using the equation: η(%) = (W − Wt)/W × 100, where W is the initial mass of the catalyst and Wt is the remaining mass after use. After five consecutive photocatalytic cycles, the ZIF-8/GO catalyst exhibited a suspension retention of only 12.8%, as shown in Fig. 4(h). This low value suggests that the photocatalytic process predominantly took place at the catalyst's surface rather than within the bulk liquid phase. To further assess the catalyst's durability, both its morphology and elemental composition were examined using transmission electron microscopy (TEM) before and after the reuse tests (Fig. 4(i)). TEM images revealed that the morphology of ZIF-8/GO remained intact even after five cycles, validating its structural stability. Additionally, EDX analysis was performed to validate the elemental composition after the stability evaluation. The analysis identified the occurrence of C 1s, O 1s, N 1s, and Zn 2p elements, as illustrated in Fig. 4(i). These results indicate that the structural morphology and the necessary elemental composition of the ZIF-8/GO were preserved before and after the stability test, confirming the robustness and stability of the structure.
Prior to the photocatalytic experiments, the degradation behavior of 5-FU under visible light irradiation without any catalyst was evaluated over a 30-minute period the corresponding absorbance spectrum is presented in Fig. 5(a). The results demonstrated that the absorbance of 5-Flu remained nearly unchanged throughout the exposure time, indicating that visible light alone was insufficient to induce photodegradation. Based on results, 5-Flu is stable and unaffected to breakdown by visible light alone, and a catalyst is necessary to facilitate its degradation.
Upon introducing photocatalysts such as GO, ZIF-8, and their composite, a notable enhancement in 5-Flu degradation under visible light was observed, emphasizing the photocatalytic potential of these materials. When GO was employed, the absorbance of 5-Flu gradually declined over a 100-minute irradiation period, as shown in Fig. 5(b). Significantly, the absorbance intensity of 5-Flu decreased more rapidly under similar time-frame, and corresponding results are shown in Fig. 5(c).
The use of ZIF-8/GO composite led to a significant decrease in the absorbance intensity of 5-Flu over 100-min visible light irradiation as depicted in Fig. 5(d). The optimal ZIF-8/GO photocatalyst demonstrated a remarkably high degradation efficiency for 5-Flu, achieving 97.4% degradation with a kinetic rate constant of 0.026 min−1. This efficiency is significantly higher compared to GO (42.1%, 0.0057 min−1) and ZIF-8 (67.7%, 0.011 min−1), as the detailed calculated results at given interval of time are depicted in Fig. 5(e). This enhanced performance can be attributed to the synergistic effects of the core–shell structure, which boosts the overall efficiency of the photocatalysis.
The recyclability of ZIF-8/GO composite was also investigated through employing six consecutive photocatalytic cycles to assess its stability. After each cycle, 2 mL of the 5-Flu suspension was extracted and centrifuged to separate the solution from the catalyst, which was then washed and reused. The catalyst achieved 97.4% degradation of 5-Flu in the first cycle. Subsequent cycles showed degradation efficiencies of 96.2%, 94.5%, 93.3%, 91% and 88.2%, and the results are shown in Fig. S7. Despite a 12.8% decrease in efficiency by the six cycles, the ZIF-8/GO composite maintained high photocatalytic activity, with a final efficiency of 88.2%, indicating good stability and reusability.
In addition, the impact of pH on the degradation of Rho-B and 5-Flu were conducted to assess how varying pH levels affect the photocatalytic efficiency of the ZIF-8/GO composite. Various experiments were collected in the range pH values, alkaline (pH 9), neutral (pH 7) and acidic (pH 3), to assess their effect on the degradation process. With Rho-B, the degradation efficiency varied with pH. Under an acidic environment (pH 3), the absorbance spectra of Rho-B showed moderate reduction after 100 min of visible light irradiation. In a neutral system (pH 7), the absorbance intensity decreased more significantly compared to the acidic condition. Overall, the ZIF-8/GO composite demonstrated good photocatalytic behavior in an alkaline medium with a pH of 9. The degradation efficiency of Rho-B was 47.3%, 64.8%, and 100% at pH values of 3, 7, and 9, when 100 min visible light exposed, respectively (Fig. 6(c)).
The highest degradation efficiency was achieved at alkaline pH (pH 9), where the catalyst performed optimally. At acidic pH (pH 3), efficiency decreased, likely due to increased protonation of the dye, which may hinder its interaction with the catalyst. n neutral conditions (pH 7), the degradation efficiency may have reduced because of suboptimal conditions for the catalyst's reactivity. At neutral pH, the catalyst might not be in its most effective state for interacting with the dye, potentially resulting in less efficient degradation compared to the alkaline environment where the catalyst could be more reactive or better suited to the conditions.
For 5-Flu, the degradation efficiency was notably also influenced by pH. The highest performance was observed at neutral pH (pH 7, 97.4%, suggesting that the catalyst functions most effectively under these conditions. Degradation efficiency decreased at acidic pH (pH 3, 66.9%) and alkaline pH (pH 9, 76.9%), which could be due to changes in reactivity or stability of the catalyst at these pH levels. The corresponding results are shown in Fig. 6(d). The observed behavior can be attributed to the following factors: at lower pH levels, the primary active species, such as ˙OH and SO4˙−, may react with excess h+ ions to form less reactive substances (e.g., H2O and HSO4−). Conversely, at higher pH levels, some SO4˙− may react with OH−, producing ˙OH with a relatively weaker oxidation potential. Therefore, a neutral pH environment is more suitable for the ZIF-8@GO catalyst, as it facilitates better degradation of 5-Flu under visible light, indicating its effective applicability in photocatalysis. These results indicate that the pH of the solution plays a crucial role in the photocatalytic degradation of both Rho-B and 5-Flu, with neutral conditions typically offering the most effective environment for optimal catalytic performance.
Under light irradiation, ZIF-8 absorbs photons and generates electron–hole pairs (reaction 1). The photogenerated electrons are transferred to the GO sheets (reaction 2), where they interact with dissolved oxygen to yield superoxide radicals (reaction 3). These radicals can further react with water to produce hydroxyl radicals (reaction 4), both of which actively degrade the Rho-B molecules (reactions 5 and 6). Meanwhile, holes in ZIF-8 may contribute to oxidation processes, as supported by the moderate influence of EDTA-2Na. The following reactions summarize the mechanistic pathway:
ZIF − 8 + hv → ZIF − 8 [e− + h+] | (1) |
e− + GO → GO [e−] | (2) |
GO [e−] + O2 → ˙O2− | (3) |
˙O2− + H2O → ˙OH + OH− | (4) |
˙OH + Rho-B → [degraded products] | (5) |
O2− + Rho-B → [degraded products] | (6) |
e− + ![]() | (7) |
h+ + EDTA-2Na → EDTA-2Na [oxidized specie] | (8) |
˙OH + FFA → FFA [oxidized] | (9) |
˙O2− + p-BQ → p-BQ [oxidized] | (10) |
The results highlight the synergistic role of the ZIF-8/GO composite in facilitating effective charge separation and promoting the generation of reactive oxygen species, leading to enhanced photocatalytic activity.
To validate the universality of this mechanism, analogous scavenger experiments were conducted using 5-fluorouracil (5-Flu) as the target pollutant under identical conditions. The observed trends closely mirrored those for Rho-B. Specifically, the addition of FFA and p-BQ significantly reduced degradation efficiency, confirming the central role of ˙OH and ˙O2− radicals. On the other hand, the effects of K2Cr2O7 and EDTA-2Na were comparatively minor, reinforcing the idea that ROS, rather than the direct action of charge carriers, drive the photocatalytic reaction. Fig. 7(b) presents the comparative degradation efficiencies of 5-Flu under the influence of each scavenger, while detailed UV-vis spectral analyses are provided in Fig. S12(a–d), further supporting the proposed mechanism. These results demonstrate that the ZIF-8/GO composite effectively promotes charge separation and facilitates the formation of ˙OH and ˙O2− radicals, which are the principal oxidants for both Rho-B and 5-Flu degradation. Overall, this study underscores the efficacy and versatility of the ZIF-8/GO photocatalyst in environmental remediation, showcasing its broad-spectrum capability in degrading structurally distinct organic pollutants via a unified ROS-driven mechanism.
The mechanism can also be rationalized through the band alignment of the ZIF-8/GO composite. Although we did not directly determine the band edges in this work due to the absence of the required facilities, the explanation can be constructed based on well-documented literature values. ZIF-8 is reported to possess a relatively wide bandgap of about 4.3–5.1 eV, with its conduction band positioned near −0.9 eV versus NHE, while GO exhibits a narrower effective bandgap of ∼2.9 eV and functions as an efficient electron acceptor owing to its conductive nature and favorable work function.62–64 Upon visible-light excitation, electrons generated in the conduction band of ZIF-8 can readily migrate to the GO sheets, which not only prevents rapid recombination with holes but also facilitates the reduction of dissolved oxygen into ˙O2− radicals. Meanwhile, the holes retained in the valence band of ZIF-8 participate in oxidation reactions, thereby sustaining the overall photocatalytic cycle.65 This band alignment thus explains the improved charge separation and transfer dynamics, ultimately leading to the higher degradation efficiency observed in the ZIF-8/GO system, consistent with prior findings for similar MOF/GO photocatalysts.66,67
Photocatalysis includes the acceleration of a photoreaction in the existence of a catalyst.68 Using this mechanism, the catalyst is generally a semiconductor material that, upon absorbing light, generates electron–hole pairs which can function chemical reactions. To further explore the photocatalytic activity mechanisms of the ZIF-8/GO composite, cyclic voltammetry (CV) and electrochemical impedance spectroscopy tests were conducted using a three-electrode setup in 3 M KOH aqueous electrolytes. The potential window range was applied from 0 to 0.5 V for CV analysis and EIS study was performed in the frequency range of 100 kHz–100 MHz. The CV electrochemical technique benefits to understand the redox method occurring on the surface of photocatalyst (ZIF-8/GO) and the contribution of numerous reactive species generated throughout the photocatalytic reaction. CV is a promising technique to probe the redox behavior of photocatalysts. As it includes sweeping the potential of an electrode engrossed in an electrolyte solution while measuring the resulting current. CV analysis of ZIF-8/GO demonstrates a well-defined redox peak and also express the excellent cycle stability, showed by overlapping traces (Fig. S13). The CV profile of ZIF-8/GO is consistent with our prior findings. As shown in figure, there are two distinct, reversible redox processes labeled reduction and oxidation for ZIF-8/GO, with the sharp increase at 0.5 V corresponding to the irreversible oxidation of water (Fig. 7(c)). Alone GO did not show any redox process, it suggests that its intrinsic photocatalytic activity is limited. This limitation can be attributed to several factors such as poor charge separation, low light absorption, and rapid recombination of photogenerated electron–hole pairs. The CV profile of alone GO is being consistent with the photodegradation of Rho-B and 5-Flu. The peak separations between the oxidation and the corresponding reduction are larger for ZIF-8 (0.15 V) compared to ZIF-8/GO (0.10 V), indicating that dye reduction occurs more slowly with ZIF-8 than with ZIF-8/GO. The faster dye reduction with ZIF-8/GO can be attributed to the following factors; (i) enhanced charge transfer: the presence of ZIF-8 in the composite may facilitate better charge transfer between GO and the Rho-B, leading to more efficient electron flow and faster reduction reactions, and (ii) improved surface interactions: the combination of ZIF-8 with GO could enhance the interaction between the ZIF-8/GO and the Rho-B, increasing the accessibility of Rho-B molecules to reactive sites on the surface of ZIF-8/GO. Moreover, at identical material loading, the ZIF-8/GO composite exhibits a significantly higher current density (5.9 mA), nearly doubling that of pure ZIF-8 (3.0 mA). This higher current density indicates that ZIF-8/GO exhibits superior catalytic activity compared to ZIF-8 and GO alone (1.0 mA).
The higher catalytic activity of ZIF-8/GO compared to ZIF-8 and GO alone, as reflected by the increased current density, can be attributed to several key factors. Firstly, the incorporation of GO into ZIF-8 significantly enhances the material's electronic conductivity. GO provides a conductive network that facilitates better electron transfer across the composite, thereby improving the overall catalytic performance. In addition, GO contributes additional active sites for catalytic reactions, which interact more effectively with reactants, further increasing the catalytic efficiency. The synergy between ZIF-8 and GO also results in optimized charge separation and reduced recombination of photogenerated electron–hole pairs, allowing more carriers to participate in catalytic processes. Moreover, the combination of ZIF-8 and GO may improve the surface properties of the material, such as its electronic structure and surface morphology, which enhances the interaction with reactants and boosts catalytic activity. However, the CV profile of ZIF-8/GO shows higher current density and smaller peak separations, suggesting that the improved photocatalytic performance is due to increased photocurrent and enhanced charge separation.
As shown in Fig. 7(d), the Nyquist plot reveals that ZIF-8/GO has a very low series resistance (Rs) of 1.4 Ω, compared to ZIF-8 (3.5 Ω) and GO (5.7 Ω). Notably, the charge transfer resistance (Rct) for ZIF-8/GO is approximately 0.2 Ω, which is significantly lower than the Rct values of 2.2 Ω for ZIF-8 and 4.1 Ω for GO. This suggests enhanced separation of photogenerated charges during the oxidation process and a higher density of active sites. The findings confirm that ZIF-8/GO facilitates more efficient charge transfer from the photocatalyst to the solution and exhibits lower charge transport resistance,69,70 highlighting its enhanced photocatalytic activity.
This journal is © The Royal Society of Chemistry 2025 |