Van Nhuong Vua,
Thi Ha Thanh Phama,
Thi Huong Lea and
Truong Xuan Vuong
*b
aFaculty of Chemistry, Thai Nguyen University of Education, No. 20 Luong Ngoc Quyen Street, Thai Nguyen City 24000, Vietnam
bFaculty of Natural Sciences and Technology, TNU-University of Science, Tan Thinh Ward, Thai Nguyen City 24000, Vietnam. E-mail: xuanvt@tnus.edu.vn
First published on 11th August 2025
This study investigates Cu(II)-modified Mg–Al layered double hydroxide (LDH) composites for efficient removal of phenol red (PR) from water through adsorption and visible-light photocatalysis. The aim was to develop a sustainable material capable of addressing dye pollution in textile wastewater. The composites were synthesized via co-precipitation with optional calcination, non-calcined and calcined materials. Structural characterization (XRD, EDX, FT-IR, SEM) confirmed successful Cu2+ incorporation by isomorphic substitution, forming mesoporous, less crystalline structures. Uncalcined samples, particularly 5-CuH, retained a layered morphology and removed up to 94% of PR. The removal occurred via electrostatic attraction, π–π interactions, and ion exchange. Cu2+ narrowed the bandgap (Eg ≈ 2.1–3.1 eV), enhancing photocatalytic activity. Under visible light, 5-CuH degraded approximately 40% of PR within 60 minutes. Calcined samples (e.g., 5-CuH500) formed CuO particles and performed better under acidic conditions despite reduced structural integrity. Both materials reduced COD by over 90% in real textile wastewater, confirming their dual-function performance, adsorption and photocatalysis. In conclusion, Cu–Mg–Al composites offer an effective, eco-friendly alternative for treating dye-contaminated wastewater.
The environmental impact does not end with ecosystems. PR accumulated in water has also been linked with severe health risks such as cancer, genetic impairment, and disease of the reproductive system.5 Thus, the treatment of colored wastewater is required to safeguard not only environmental health but also public health.
Of the available treatments, adsorption is in general reputed to be easy to apply, low-cost, and very effective at removal, even at low concentrations of pollutants.6 Activated carbon, clays, and layered double hydroxides (LDHs) are typical adsorbents.7–9 However, issues still remain regarding optimizing material efficiency without sacrificing economic scalability.10 Activated carbon and clays are widely used adsorbents but each has notable limitations. Activated carbon, while effective, is costly to produce, difficult to regenerate, and lacks selectivity.11,12 Clays typically have low adsorption capacity, slow kinetics, and are sensitive to pH changes, often aggregating in water.13
Photocatalytic degradation, another effective process, uses light to activate catalysts that generate reactive oxygen species (ROS), which break down organic dyes.14 The process is very efficient under sunlight.15 Weak visible-light absorption, however, is a drawback that reduces the efficiency of most photocatalysts.16 Therefore, scientists have looked for advanced materials, including heterojunctions and polyoxometalate-based systems, to increase efficiency and broaden applicability.17,18
To overcome the limitations of a single method, a hybrid process integrating adsorption and photocatalysis in an integrated material system has been achieved. The approach utilizes the pre-concentration of pollutants through adsorption to enhance the efficiency of photocatalytic degradation.19 LDHs, particularly Zn–Al and Mg–Al types, are highly promising candidates due to their adjustable structures, high surface areas, and multi-functionality.9,20,21
Recent studies have shown that doping Zn–Al LDHs with transition metal ions like Cu2+ greatly increases photocatalytic activity, primarily by enhancing charge separation and improving visible-light response.9 Doping by Cu2+ also augments adsorption, ensuring dual-functionality for dye cleaning.22 Synergistic interaction between photocatalysis and adsorption creates more active sites and facilitates efficient degradation channels. Cu2+-modified Mg–Al layered double hydroxides (LDHs) have been synthesized and utilized as efficient catalysts for the synthesis of organic compounds,23–25 and the removal of methyl orange from water23 owing to their unique structural and physicochemical properties. Despite such advancements, the application of Cu2+-modified Mg–Al LDHs towards the removal of phenol red has not been adequately explored. The mechanism behind Cu2+-promoted photocatalysis is not yet fully understood.9,26 To address this knowledge gap, the present work develops a Cu2+-modified Mg–Al LDH system that integrates adsorption and photocatalysis to realize the effective and sustainable removal of phenol red from wastewater.
The present study focuses on the synthesis and evaluation of Cu2+-modified Mg–Al layered double hydroxides (LDHs) for the effective removal of phenol red (PR) from aqueous solutions. The proposed Cu(II)-modified LDH composite is designed to serve as a multifunctional material with several desirable properties: (i) high adsorption capacity for PR due to its well-ordered layered structure; (ii) the ability to initiate photocatalytic degradation of PR under visible light irradiation; (iii) strong structural stability, enabling repeated use without significant performance loss; and (iv) efficient operation under near-neutral pH conditions, enhancing its applicability in practical water treatment systems. To the best of our knowledge, this is the first report on the synthesis of a Cu(II)-modified Mg–Al LDH composite specifically tailored for the efficient removal of phenol red from water.
This research pursues three main objectives: (1) to investigate the adsorption performance of the modified LDHs; (2) to evaluate their photocatalytic activity under visible light; and (3) to examine the synergistic effect of combining adsorption and photocatalysis.
We hypothesize that Cu2+ modification enhances the material's performance through increased surface area, improved charge separation, and the inherent catalytic activity of Cu2+ under light activation. By integrating adsorption and photocatalysis into a single system, this approach offers a more efficient and sustainable method for the removal of phenol red from water.
The hydrotalcite and Cu-hydrotalcite samples were subsequently calcined at 500 °C for 5 hours with a heating rate of 3 °C min−1, and were labeled as H500 and n-CuH500, respectively. A detailed description of the sample nomenclature is provided in Table 1.
Sample | Molar ratio of elements Cu![]() ![]() ![]() ![]() ![]() ![]() |
---|---|
H (H500) | 0![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
2-CuH (2-CuH500) | 0.02![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
3-CuH (3-CuH500) | 0.03![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
4-CuH (4-CuH500) | 0.04![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
5-CuH (5-CuH500) | 0.05![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
6-CuH (6-CuH500) | 0.06![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
The samples were then ground using a mortar and pestle and were employed for the characterization of their structural properties and photocatalytic activity in the degradation of phenol red and dye pollutants in textile wastewater.
![]() | (1) |
where: C0 is the initial concentration of phenol red, C is the concentration of phenol red at the time of sampling.
The adsorption capacity of 12 synthesized material samples (6 non-calcined and 6 calcined at 500 °C) was investigated. All experiments were repeated three times. From the results, the adsorption efficiency and the equilibrium adsorption time for the materials toward phenol red in aqueous solution were determined.
To assess the phenol red degradation capacity of the material series over time, samples were taken at fixed intervals (every 30 minutes). Afterward, the samples were centrifuged, diluted five times, and analyzed using a UV-Vis 1700 spectrophotometer. From the obtained results, the photocatalytic degradation efficiency was determined using the following eqn (2)
![]() | (2) |
where: Ccb: concentration of phenol red (PR) at adsorption equilibrium, C: concentration of phenol red (PR) at the time of sampling.
However, both adsorption and photocatalytic degradation processes conducted simultaneously during the treatment of phenol red. Therefore, the degradation efficiency of phenol red can be determined according to the eqn (3):
Photocatalytic degradation efficiency PR (%) = treatment efficiency PR (%) − adsorption efficiency PR at adsorption equilibrium (%) | (3) |
The treatment efficiency of phenol red (including adsorption and photocatalytic degradation) was determined according to eqn (4):
![]() | (4) |
where: C0 is the initial concentration of phenol red, C is the concentration of phenol red at the time of sampling.
The degradation capacity of the 12 synthesized material samples was investigated. All experiments were repeated three times. From the obtained results, the phenol red degradation efficiency of the synthesized material samples over time and in relation to the Cu2+ ion ratio in the materials was evaluated.
The experimental procedure was carried out in the same manner as described above, simultaneously for all five concentrations of phenol red (PR). All experiments were repeated three times. From the obtained results, we were able to assess the effect of PR concentration on the catalytic activity of the material.9
A 250 mL diluted wastewater sample was placed in a 500 mL beaker, and the pH of the solution was adjusted to different values within the optimal pH range. Then, 0.2 g of the optimal material sample was added, and the beaker was sealed with a black plastic bag. The mixture was stirred on a magnetic stirrer for 150 minutes at room temperature to achieve adsorption equilibrium of the material. Subsequently, 5 mL of 30% H2O2 was added to the beaker, which was again sealed with the black plastic bag and illuminated under a 30 W LED light source at room temperature.9 Samples were taken every 60 minutes to determine the degradation of colorants and the mineralization of organic compounds in the wastewater for both non-calcined and calcined synthesized material samples.9
The degradation efficiency of colorants in the wastewater sample is determined using the following formula:
![]() | (5) |
The mineralization efficiency of COD in the wastewater is determined using the following formula:
![]() | (6) |
Chemical Oxygen Demand (COD) is determined using the K2Cr2O7 method, with detailed procedures provided in the previous studies.9
Table 2 illustrates the d003 values, representing the interlayer spacing. For uncalcined MgAlCO3 and CuMgAlCO3 samples, d003 values range from 7.66 to 8.03 Å, which is characteristic of hydrotalcite-like structures and confirms the presence of interlayer CO32− anions.27 Additionally, the weakening of the characteristic hydrotalcite diffraction peaks can be attributed to the isomorphic substitution of Mg2+ by Cu2+ ions and the slightly larger ionic radius of Cu2+ (0.073 nm) compared to that of Mg2+ (0.072 nm).24 Although some of the uncalcined samples appear black in colour, no diffraction peaks corresponding to crystalline CuO, such as the one at 2θ = 35.4° (d200), are observed.24 This indicates that Cu2+ ions are well-dispersed within the brucite-like layers of the MgAlCO3 hydrotalcite structure.
Sample | d003 (Å) | d006(Å) | d110(Å) | Average crystallite size (Scherrer method) (nm) | Lattice parameters (Å) | Color | |
---|---|---|---|---|---|---|---|
a | c | ||||||
a “—”: undetermined. | |||||||
H | 7.870 | 3.819 | 1.524 | 7.02 | 3.048 | 23.601 | White |
2-CuH | 8.030 | 3.887 | 1.525 | 4.58 | 3.050 | 24.090 | Black |
3-CuH | 7.843 | 3.852 | 1.521 | 4.17 | 3.042 | 23.880 | Black |
4-CuH | 7.843 | 3.826 | 1.522 | 4.20 | 3.044 | 23.529 | Black |
5-CuH | 7.807 | 3.809 | 1.521 | 3.43 | 3.042 | 23.529 | Blue + black |
6-CuH | 7.664 | 3.869 | 1.52 | 3.08 | 3.040 | 23.421 | Blue |
H500 | 7.704 | 3.816 | 1.519 | 6.96 | 3.038 | 22.992 | White |
2-CuH500 | — | — | — | 10.24 | — | — | Black |
3-CuH500 | — | — | — | 9.27 | — | — | Black |
4-CuH500 | — | — | — | 13.49 | — | — | Black |
5-CuH500 | — | — | — | 19.56 | — | — | Black |
6-CuH500 | — | — | — | 25.49 | — | — | Black |
Table 2 presents the calculated values of the lattice parameters a and c. The parameter a represents the distance between cations within the brucite-like layer and is calculated using the formula a = 2·d110. The parameter c corresponds to the thickness of the brucite layer combined with the interlayer spacing and is determined by c = 3·d003.22,28 The a values range from 3.04 to 3.05 Å, while the c values lie between 23.0 and 24.09 Å (Table 2). These parameters vary only slightly with the increase of Cu2+ content in the modified materials. These results indicated the preservation of the hydrotalcite-like layered structure and the isomorphic substitution of Mg2+ by Cu2+ ions within the brucite-like lattice.
The XRD analysis results of Cu2+-modified hydrotalcite materials are shown in Fig. 1b. After calcination at 500 °C for 5 hours, the synthesized samples exhibit significant changes in both characteristic diffraction peaks and crystalline phase composition. Diffraction peaks corresponding to the hydrotalcite structure are still observed in the H500 sample after calcination, appearing at angles associated with the (003), (006), (009), and (110) planes. This indicates that the H500 sample retains its hydrotalcite-like layered structure post-calcination, although the peak intensities are significantly reduced.
In contrast, the characteristic peaks of the hydrotalcite structure completely disappear in the Cu2+-modified samples after calcination. This is due to the complete decomposition of the materials into oxides such as MgO, Al2O3, and CuO.29,30 However, no diffraction peaks corresponding to the MgO phase are observed. Only two prominent peaks of the CuO phase appear at 2θ angles of 35.4° and 38.6°, with their intensities increasing progressively with the rising Cu/Al molar ratio in the samples.
The average crystallite sizes of CuO are provided in Table 2. The CuO crystallites are in the nanometer range, with sizes ranging from 9.3 to 25.5 nm. The increase in CuO crystallite size may be attributed to the aggregation of smaller particles, likely caused by the increased Cu2+ content in the synthesized materials.
Sample | Atomic composition of elements in studied samples (%) | |||
---|---|---|---|---|
% Mg | % Cu | % Al | % O | |
H | 17.25 | 0 | 7.57 | 75.18 |
H500 | 21.35 | 0 | 8.88 | 69.77 |
2-CuH | 14.15 | 5.14 | 8.11 | 72.60 |
2-CuH500 | 17.00 | 6.67 | 9.93 | 66.40 |
5-CuH | 5.76 | 14.74 | 8.82 | 70.68 |
5-CuH500 | 7.97 | 19.82 | 11.18 | 61.02 |
The results indicate that all six samples, including three uncalcined (H, 2-CuH, 5-CuH) and three calcined (H500, 2-CuH500, 5-CuH500), contain the elements Mg, Cu, Al, and O. The atomic ratios of elements such as Mg:
Al and Cu
:
Mg
:
Al in the analyzed samples closely match the theoretical ratios calculated for their synthesis. For example, sample H shows a Mg
:
Al ratio of 17.25
:
7.57 ≈ 2.28
:
1, which corresponds well with the theoretical ratio of 7
:
3. Sample 5-CuH exhibits a Cu
:
Mg
:
Al ratio of 14.74
:
5.76
:
8.82 ≈ 5.12
:
2
:
3.06, approximating the designed ratio of 5
:
2
:
3.
These EDX results confirm that the metal salt precursors were almost completely precipitated as hydroxides, which then formed the hydrotalcite and Cu2+-modified hydrotalcite structures.
The sharp bands at 1383.31–1383.98 cm−1 are characteristic of the stretching vibrations of interlayer carbonate (CO32−) anions. In the lower wavenumber region (448.02–834.36 cm−1), the observed bands are associated with metal–oxygen (M–O), metal–hydroxyl (M–OH), and bridging oxygen vibrations (O–M–O, M–O–M), indicating the presence of Cu2+, Mg2+, and Al3+ cations within the brucite-like layers. Both 5-CuH and 5-CuH500 samples exhibit additional peaks at 834.36 and 803.20 cm−1, which are assigned to Cu2+ species incorporated into the hydrotalcite and LDH structures. These results confirm the presence of interlayer carbonate anions and the successful incorporation of Cu2+ ions into the brucite-like lattice of the synthesized materials.9,31
After calcination, the H500 samples had structural changes due to the decomposition of hydroxide groups and interlayer carbonates. Although the layered hydrotalcite-like structure was partially retained, many sheets were broken into smaller fragments or aggregated into larger ones, resulting in less uniform morphology compared to the uncalcined H sample.
In contrast, the 5-CuH500 sample exhibited the destruction of the hydrotalcite-like layered structure. The SEM image of 5-CuH500 revealed the presence of small nanoparticles (tens of nanometers) and larger aggregates (>100 nm), with no visible layered morphology. These SEM observations were in good agreement with the XRD analysis results, confirming the structural changes induced by Cu2+ incorporation and thermal treatment.
Sample | BET surface area (m2 g−1) | Pore diameter (nm) | Pore volume (cm3 g−1) |
---|---|---|---|
H | 67.76 | 27.6 | 0.542 |
2-CuH | 57.57 | 27.9 | 0.402 |
5-CuH | 41.39 | 23.4 | 0.242 |
H500 | 58.88 | 26.4 | 0.389 |
2-CuH500 | 50.05 | 27.5 | 0.345 |
5-CuH500 | 26.34 | 16.7 | 0.109 |
![]() | ||
Fig. 5 Nitrogen adsorption–desorption isotherms of the synthesized materials of noncalcined materials (a) and calcined materials (b). |
The nitrogen adsorption–desorption isotherms (BET) presented in Fig. 6 reveal that all six synthesized materials exhibit type IV isotherms with H3 hysteresis loops according to the IUPAC classification, which is characteristic of mesoporous materials.9,31 The narrow hysteresis loops observed in the isotherms indicate that the BET surface area and pore volume of the analyzed materials are relatively low. In particular, the 5-CuH500 sample shows a more extended nitrogen adsorption–desorption curve over the entire relative pressure range (P/P0 from 0 to 1.0), suggesting the coexistence of both mesoporous and microporous structures in this sample.
The data presented in Table 3 and Fig. S1 (SI) showed that both the BET surface area and pore volume of the non-calcined samples (H, 2-CuH, and 5-CuH) and the calcined samples (H500, 2-CuH500, and 5-CuH500) decreased significantly as the Cu/Al molar ratio increases from 0 to 5/3. Additionally, the average pore diameter also tended to decrease with increasing Cu content. This reduction in surface area and pore volume was attributed to the deterioration of the hydrotalcite-like layered structure due to Cu2+ incorporation, leading to partial or complete structural collapse, particularly after thermal treatment at high temperatures. These findings are consistent with the XRD and SEM analyses and agree with the results reported in previous studies.22,26
Fig. 6C illustrates the UV-Vis DRS spectra of samples calcined at 500 °C. After calcination, the H500 sample also displays two major absorption peaks at around 220 and 320 nm, similar to the uncalcined H sample. However, the peak intensity at 320 nm in the H500 sample is significantly lower than that of the uncalcined H. In contrast, the calcined CuH samples exhibit a marked red shift of the absorption edge into the visible region. The UV-Vis DRS spectra of calcined samples differ markedly from those of uncalcined ones. This difference can be attributed to the formation of CuO nanoparticles in the materials after calcination, which strongly affects the absorption edge and significantly reduces the band gap energy.
Fig. 6B and D show the Tauc plots used to determine the band gap energy of the materials. The H and CuH samples follow an indirect allowed transition model (Tauc method with exponent m = 1/2), whereas the calcined H500 and CuH500 samples correspond to a direct allowed transition (m = 2).35 Based on these results, the band gap energy (Eg) of the synthesized materials was calculated and summarized in Table S1 (see SI). A significant decrease in Eg was observed with increasing Cu2+ ion content, suggesting that Cu2+-modified materials are likely to exhibit enhanced photocatalytic activity under visible light irradiation.
![]() | ||
Fig. 7 Adsorption efficiency of 100 ppm phenol red (PR) over time on non-calcined materials H and n-CuH (n = 2–6). |
![]() | ||
Fig. 8 Speciation of phenol red in different environments.36 |
Electrostatic attraction between PR anions and metal cations on the brucite-like layers, as well as anion exchange with interlayer carbonate (CO32−), likely facilitated the adsorption. A decrease in BET surface area may have further influenced the efficiency. The MgAlCO3 and CuMgAlCO3 materials share similar structures with ZnAlCO3 and CuZnAlCO3 and also adsorbed PR effectively. However, some MgAlCO3- and CuMgAlCO3-based samples achieved higher PR adsorption efficiency than those in the Zn-based series.9 Their greater BET surface areas likely contributed to this improved performance.
![]() | ||
Fig. 9 Adsorption efficiency of 100 ppm phenol red (PR) over time on calcined materials H500 and n-CuH500 (n = 2–6). |
In addition to factors such as the layered hydrotalcite structure, BET surface area, pore volume, pore diameter, and electrostatic interactions between PR anions and metal cations, the so-called memory effect played a significant role in enhancing the adsorption performance of calcined samples such as H500, 2-CuH500, and 3-CuH500.26,37 Based on the evaluation of PR (100 ppm) adsorption under dark conditions, we further investigated the photocatalytic activity of the synthesized materials by assessing their ability to degrade PR. The study also identified optimal conditions for PR degradation, including illumination time, Cu2+ content in the material, PR concentration, and solution pH, and evaluated the materials' potential for application in the treatment of textile dyeing wastewater.
In addition, surface hydroxyl groups, interlayer anions, and the modified Cu2+ sites collectively contribute to the adsorption process through ion exchange, hydrogen bonding, and coordination interactions.9 These synergistic mechanisms significantly boost the adsorption capacity and efficiency of the Cu2+-modified materials compared to unmodified Mg–Al LDHs.
![]() | ||
Fig. 11 The treatment efficiency of 100 ppm phenol red (PR) over time on calcined samples H500 and n-CuH500 (n = 2–6). |
However, the H sample had a large bandgap energy (Eg) and mainly absorbed UV light, resulting in negligible PR (100 ppm) degradation. In contrast, Cu2+-modified MgAlCO3 samples exhibited excellent photocatalytic performance. Among the non-calcined modified materials, 5-CuH showed the highest photocatalytic activity, achieving a total treatment efficiency of approximate 94% within 60 minutes of illumination, including about 40% from photocatalytic degradation. Based on Table S4 and Fig. 10, the photocatalytic activity decreased in the order: 5-CuH > 6-CuH > 4-CuH > 3-CuH > 2-CuH > H. The high activity under 30 W LED light is attributed to the synergy between appropriate bandgap energy, high BET surface area, Cu2+ active sites, illumination time, and the presence of H2O2 as an oxidizing agent. The PR degradation mechanism can be explained by the following reactions.9,22,39
Cu2+ − MgAl + hυ → Cu2+ − MgAl (e−, h+) | (7) |
h+ + H2O2 → 2˙OH | (8) |
H2O2 + e− → ˙OH + OH− | (9) |
OH˙ + PR → colorless reduction products | (10) |
h+ + PR → colorless oxidation products | (11) |
Compared to CuZnAlCO3 materials with Cu/Al molar ratios of 3/3 and 3.5/3 (samples CuH-3.0 and CuH-3.5), the 5-CuH sample exhibited significantly higher photocatalytic activity. This suggests that a Cu/Al molar ratio of 5/3 may be optimal for obtaining materials with superior photocatalytic performance.9
Compared to the calcined samples, the uncalcined samples 4-CuH, 5-CuH, and 6-CuH demonstrated higher PR (100 ppm) treatment efficiency than their respective calcined counterparts (4-CuH500, 5-CuH500, and 6-CuH500) under identical conditions and with equal Cu2+ content. This result indicates that the layered hydrotalcite structure and the incorporation of Cu2+ ions into the brucite-like lattice play a crucial role in enhancing the photocatalytic activity of uncalcined materials. Based on these findings, the two representative samples, 5-CuH (uncalcined) and 5-CuH500 (calcined) were selected to further investigate the effects of PR concentration and pH on photocatalytic performance, as well as their potential for treating real textile dyeing wastewater (mat weaving effluent).
![]() | ||
Fig. 12 Graph illustrating the treatment efficiency of PR at different concentrations using the 5-CuH material. |
![]() | ||
Fig. 13 The treatment efficiency of phenol red (PR) at various concentrations using the 5-CuH500 material. |
The results presented in Table S6 and Fig. 12 showed that the treatment efficiency of PR gradually decreased as the initial PR concentration increased from 100 to 200 ppm. At a concentration of 100 ppm, the treatment efficiency exceeded 90% after just 60 minutes of illumination. However, for higher concentrations (150, 175, and 200 ppm), the illumination time required to reach approximately 83% treatment efficiency extended to 240 minutes. Despite the reduced efficiency at higher concentrations, the 5-CuH material still demonstrated a strong photocatalytic capacity, capable of degrading PR at concentrations up to 200 ppm. This performance surpasses that of previously reported CuH-3.0 and CuH-3.5 materials,9 highlighting the superior photocatalytic activity of 5-CuH and its potential for treating wastewater containing high concentrations of phenol red.
For the calcined sample 5-CuH500, the results presented in Table S7 and Fig. 13 revealed a similar degradation trend to that of the uncalcined 5-CuH sample and aligned with the general rule that increasing the pollutant concentration leads to a decrease in reaction rate and overall degradation efficiency.
When comparing PR degradation efficiencies at corresponding concentrations between the two samples containing the same amount of Cu2+ ions (5-CuH and 5-CuH500), the photocatalytic activity of 5-CuH was significantly higher than that of 5-CuH500 under identical experimental conditions. This further confirms the crucial role of the layered structure, the coordination of Cu2+ ions with O2−, and the isomorphic substitution of Cu2+ for Mg2+ in the brucite-like lattice in enhancing the photocatalytic performance of the material.
After investigating the influence of pH on the catalytic activity of the material (Section 2.4.4), we examined the treatment efficiency of phenol red at various initial concentrations using the 5-CuH500 sample at pH 3.0. The results are presented in Table S8 (SI) and Fig. 14.
Results presented in Table S8 and Fig. 14 reveal that the 5-CuH500 sample exhibited significantly enhanced photocatalytic activity under acidic conditions (pH = 3.0). When comparing the degradation of PR at 100 ppm under two different pH conditions (initial pH = 4.15 and pH = 3.0), the sample showed markedly improved performance at lower pH. At pH 3.0, the material achieved over 91.3% treatment efficiency for 100 ppm PR within just 90 minutes of LED light exposure. The 5-CuH500 material also demonstrated strong photocatalytic performance at high pollutant concentrations. It degraded 200 ppm of phenol red with an efficiency of 82.5% after 180 minutes of illumination.
These findings confirm that both 5-CuH and 5-CuH500 can effectively degrade high concentrations of phenol red (up to 200 ppm) under mild experimental conditions (30 W LED light, room temperature and pressure). The excellent catalytic activity of both materials highlights their strong potential for practical wastewater treatment applications.
Lowering the pH below 4.15 resulted in a gradual decline in degradation efficiency. Although 5-CuH maintained high activity at pH 3.5, a strong acidic environment (pH 2.0) significantly reduced its efficiency due to structural degradation of the material under highly acidic conditions (pH < 3.0).9
As the pH increased from 4.15 to 6.0, 8.0, and 10.0, the catalytic activity initially improved but then declined. This decrease at higher pH values is likely due to competition between PR anions and hydroxide ions (OH−) for adsorption sites, along with increased solution viscosity, which hinders catalytic performance. These findings align well with previous studies.9
![]() | ||
Fig. 16 The treatment efficiency of 100 ppm phenol red over time using 5-CuH500 at different pH values. |
In the pH range of 4.15 to 10.0, photocatalytic activity also showed a downward trend with increasing pH. However, reducing the pH from 4.15 to 3.5, 3.0, and 2.5 significantly enhanced the material's catalytic performance. Within the optimal range of pH 2.5–3.5, the treatment efficiency of 100 ppm PR reached approximately 97% within just 90 minutes of LED light irradiation at pH 2.5.
This enhanced activity may result from the partial dissolution of CuO, MgO, and Al2O3 under weakly acidic conditions, releasing metal ions such as Cu2+ that re-adsorb onto the material surface. These ions form Cu2+-CuMgAlCO3 catalytic centers, which boost the material's ability to convert H2O2 under LED light into more hydroxyl radicals (HO˙), thereby accelerating PR degradation.
For the 6-CuH material, the experimental results clearly indicated that hydroxyl radicals (˙OH) played the dominant role in the photocatalytic degradation of PR (Fig. 17A). In contrast, for the 5-CuH500 sample, both hydroxyl radicals (˙OH) and photogenerated holes (h+) were found to contribute significantly to the degradation process (Fig. 17B). These findings are consistent with previously reported studies40,41
![]() | ||
Fig. 17 Degradation efficiency of 100 ppm phenol red (PR) using 6-CuH (A), and 5-CuH500 in the presence and absence of active species and hole scavengers (B). |
The UV-Vis spectra of PR presented in Fig. S2C and D (SI) illustrate that the intensity of the characteristic absorption peak at 435 nm decreased more slowly over time when isopropyl alcohol was added to the reaction system. Under these conditions, the degradation efficiencies of 100 ppm PR after 180 and 120 minutes of irradiation were reduced to 54.8% and 45.5% for 6-CuH and 5-CuH500, respectively (Fig. 17A and B). These results confirm that isopropyl alcohol effectively inhibits the activity of hydroxyl radicals (˙OH), thereby slowing down the photocatalytic degradation of PR for both materials.
The results indicated that PR degradation efficiency remained low in cases (a), (b), and (d) (Fig. 18A and B). Specifically, in the condition where only H2O2 and light were applied, the degradation efficiency of 100 ppm PR reached only 9.3% at pH 4.15 and approximately 10% at pH 3.0. This suggests that the 30 W LED light source alone could activate H2O2 to a limited extent, producing only a small quantity of ˙OH radicals and resulting in poor PR degradation. Similarly low efficiencies were observed when only the catalyst was used under light or in the dark.
![]() | ||
Fig. 18 Degradation efficiency of 100 ppm phenol red (PR) using 6-CuH (A) and 5-CuH500 (B), respectively, under five different conditions. |
In contrast, when both the catalyst and H2O2 were present, the PR degradation efficiency significantly increased over time. Notably, the combination of catalyst, H2O2, and LED light irradiation produced substantially higher degradation efficiencies compared to the same system under dark conditions. This confirms that the synergistic interaction of all three components, photocatalyst, H2O2, and light, leads to a more rapid and abundant generation of hydroxyl radicals. Moreover, in the absence of light, the combined effect of the catalyst and H2O2 suggests a heterogenous Fenton-like reaction mechanism contributing to ˙OH formation.
These findings demonstrate that the simultaneous presence of catalyst, H2O2, and light markedly enhances the degradation of 100 ppm PR. Under optimized conditions, the degradation efficiency reached approximately 83% after 330 minutes using 6-CuH at pH 4.15, and about 93% after 180 minutes using 5-CuH500 at pH 3.0.
The UV-Vis absorption spectra of PR solutions (Fig. S3C, D and E (SI)) showed a characteristic peak at 435 nm, with only a gradual decrease in intensity over time when using H2O2 alone without catalyst (Fig. S3C), or in the presence of only the catalyst under light (Fig. S3D and E (SI)). These results emphasize the crucial role of H2O2 in facilitating PR degradation when used in combination with the synthesized materials.
Time (min) | The treatment efficiency of dyes in wastewater for sample 5-CuH (%) at different pH levels | |||||||
---|---|---|---|---|---|---|---|---|
pH = 3.06 | pH = 4.06 (initial) | pH = 5.5 | pH = 6.5 | |||||
Abs | H% | Abs | H% | Abs | H% | Abs | H% | |
a ad: adsorption; pH (initial): pH value of the original (untreated) dye wastewater. | ||||||||
0 | 1.715 | 0 | 1.715 | 0 | 1.715 | 0 | 1.715 | 0 |
150 ad | 1.568 | 8.6 | 0.904 | 47.3 | 0.878 | 48.8 | 1.339 | 21.9 |
180 | 0.6 | 65.0 | 0.535 | 68.8 | 0.572 | 66.6 | 0.504 | 70.6 |
210 | 0.577 | 66.4 | 0.487 | 71.6 | 0.234 | 86.4 | 0.145 | 91.5 |
270 | 0.036 | 97.9 | 0.0474 | 97.2 | 0.0301 | 98.2 | 0.0361 | 97.9 |
330 | 0.0126 | 99.3 | 0.0226 | 98.7 | 0.0159 | 99.1 | 0.0431 | 97.5 |
390 | 0.0107 | 99.4 | 0.0191 | 98.9 | 0.0281 | 98.4 | 0.023 | 98.7 |
450 | 0.0113 | 99.3 | 0.0139 | 99.2 | 0.0174 | 99.0 | 0.0264 | 98.5 |
510 | 0.0047 | 99.7 | 0.0162 | 99.1 | 0.0173 | 99.0 | 0.0289 | 98.3 |
570 | 0.0056 | 99.7 | 0.0126 | 99.3 | 0.018 | 99.0 | 0.033 | 98.1 |
630 | 0.0074 | 99.6 | 0.0166 | 99.0 | 0.0195 | 98.9 | 0.0383 | 97.8 |
Time (min) | The treatment efficiency of dyes in wastewater by the 5-CuH500 material (%) | |||||||
---|---|---|---|---|---|---|---|---|
pH = 2.5 | pH = 3.0 | pH = 3.5 | pH = 4.07 (initial) | |||||
Abs | H% | Abs | H% | Abs | H% | Abs | H% | |
a ad: adsorption; pH (initial): pH value of the original (untreated) dye wastewater. | ||||||||
0 | 1.715 | 0.0 | 1.715 | 0.0 | 1.715 | 0.0 | 1.715 | 0.0 |
150 ad | 1.707 | 0.5 | 1.676 | 2.3 | 1.318 | 23.1 | 0.851 | 50.4 |
180 | 0.614 | 64.2 | 0.724 | 57.8 | 0.682 | 60.2 | 0.503 | 70.7 |
210 | 0.577 | 66.4 | 0.66 | 61.5 | 0.614 | 64.2 | 0.438 | 74.5 |
270 | 0.504 | 70.6 | 0.608 | 64.5 | 0.492 | 71.3 | 0.099 | 94.2 |
330 | 0.164 | 90.4 | 0.346 | 79.8 | 0.079 | 95.4 | 0.055 | 96.8 |
390 | 0.0321 | 98.1 | 0.0301 | 98.2 | 0.0166 | 99.0 | 0.0279 | 98.4 |
450 | 0.0207 | 98.8 | 0.0453 | 97.4 | 0.0232 | 98.6 | 0.0492 | 97.1 |
510 | 0.0076 | 99.6 | 0.0111 | 99.4 | 0.0126 | 99.3 | 0.0197 | 98.9 |
![]() | ||
Fig. 19 Time-dependent treatment efficiency of dye compounds in diluted mat-weaving textile wastewater treated with the 5-CuH material in the presence of 30% H2O2 under visible light irradiation. |
![]() | ||
Fig. 20 Time-dependent treatment efficiency of dye compounds in diluted mat-weaving textile wastewater treated with the 5-CuH500 material in the presence of 30% H2O2 under visible light irradiation. |
Based on the obtained results, both 5-CuH and 5-CuH500 materials exhibited strong capabilities in degrading colorants present in rush mat dyeing wastewater. The 5-CuH sample achieved over 97% treatment efficiency within just 90 minutes of light irradiation across the investigated pH ranges. In comparison, the 5-CuH500 sample also reached degradation efficiencies exceeding 90%, albeit within a slightly longer irradiation period of 120 to 180 minutes.
These findings clearly confirm that both materials possess effective photocatalytic activity toward the treatment of organic colorants in dyeing wastewater. Their high performance highlights the practical applicability of these synthesized materials in the removal of industrial textile effluents.
The UV-Vis spectra of colorants in the dyeing wastewater varied over time during the adsorption and photocatalytic degradation processes, as shown in Fig. S4. At pH 6.5, the 5-CuH sample exhibited low adsorption efficiency, reaching only about 22% after 150 minutes. In contrast, the 5-CuH500 material achieved approximately 50% under the same conditions. Consequently, the intensity of the maximum absorption peak at 552 nm after 150 minutes was significantly lower for 5-CuH500 than for 5-CuH. A broad and intense absorption peak appeared at around 447 nm at the initial stage of the wastewater. However, this peak was rapidly degraded within 30 minutes of light irradiation on both materials. The absorption peak at 552 nm also declined sharply after approximately 120 minutes. Visually, the color of the solution significantly faded following centrifugation to remove solids. To accurately evaluate the mineralization of organic compounds, the samples were further irradiated for up to 10 hours. COD measurements were subsequently performed to assess the mineralization efficiency of the 5-CuH and 5-CuH500 materials.
![]() | ||
Fig. 21 Mineralization efficiency of organic compounds in mat-dyeing wastewater treated with 5-CuH (a) and 5-CuH500 (b). |
The obtained results show that a strong correlation was observed between the decrease in dye absorbance and the reduction in COD values. As dye absorbance intensity declined, COD levels also decreased. However, the COD reduction occurred at a significantly slower rate than the decrease in absorbance. After approximately 120 minutes of irradiation, dye degradation efficiencies exceeded 90% for both 5-CuH and 5-CuH500 materials. In contrast, the mineralization efficiency of organic compounds reached only 67–75% for 5-CuH and 58–69% for 5-CuH500.
These findings indicate that 5-CuH500 exhibited lower mineralization performance at the respective pH levels compared to 5-CuH. Moreover, after 10 hours of removal, 5-CuH reduced the COD of the dyeing wastewater from 486 to 49.3 mg L−1, meeting column A of the industrial wastewater discharge standard. In comparison, 5-CuH500 showed a slower mineralization rate, reducing COD from 486 to 76 mg L−1 after 12.5 hours.
Overall, both 5-CuH and 5-CuH500 demonstrated high efficiency in degrading dye compounds and effectively mineralizing refractory organics in mat-dyeing wastewater. These results suggest that the tested materials offer promising potential for advanced textile wastewater treatment applications.
Compared with state-of-the-art persulfate-activated catalysts such as CoFe2O4, although the latter may be pushed to rapid degradation rates (greater than 95% in 20 minutes), it does so at the cost of greater operational cost and complexity from the need to utilize external oxidants.42 Cu–Mg–Al LDHs, on the other hand, offer a greener, self-activating alternative that integrates both adsorption and in situ catalytic degradation, demonstrating their effectiveness and affordability for practical application.43
Conclusively, Cu2+-modified Mg–Al LDHs outperform conventional and state-of-the-art adsorbents in pollutant treatment by a large margin. Their superior adsorption capacity, kinetics, stability across a broad pH range, and decent reusability render them highly prospective for wastewater treatment applications at the practical level. Additionally, their dual-mode treatment mechanism not only enhances their efficiency but also preserves cost-effectiveness and environmental sustainability relative to more complex systems.44,45
The treatment efficiencies of 100 ppm PR solution after four successive reuse cycles of each material are presented in Fig. 22A and B. For the 6-CuH sample, a noticeable decrease in adsorption efficiency was observed after the first use, dropping sharply from 22.7% to 5.2% after 150 minutes of dark stirring. In subsequent cycles, the adsorption efficiency remained relatively stable, with only a slight decrease from 5.2% to 4.5%. Interestingly, the treatment efficiency of PR showed a modest increase over time, rising from 85.6% in the first cycle to 93.7% in the fourth cycle after 180 minutes of light irradiation.
In contrast, the 5-CuH500 material exhibited a marked decline in both adsorption capacity and photocatalytic activity across four reuse cycles. Initially, the material achieved an adsorption efficiency of 19.9% after 60 minutes of dark stirring, which significantly decreased to 1.6%, 1.9%, and 1.9% in the second, third, and fourth cycles, respectively. Under optimal conditions at pH 3.0, 5-CuH500 demonstrated a high treatment efficiency of 93.2% within just 120 minutes of illumination during the first use. However, this efficiency dropped substantially during the subsequent cycles at both 60 and 120 minutes of irradiation. When the irradiation time was extended to 180 minutes, the treatment efficiencies slightly recovered to 88.6%, 81.7%, and 81.2% for the second, third, and fourth uses, respectively. These findings are supported by UV-Vis absorption spectra of PR solutions shown in Fig. 22C and 22F
The observed reduction in adsorption and photocatalytic performance for both materials can be attributed to the gradual leaching of Cu2+ ions into the solution over time. The dissolution of copper ions likely results in a reduced number of catalytic active sites, consequently diminishing the generation of hydroxyl radicals (˙OH), which are essential for effective photocatalytic degradation. Furthermore, Cu2+ leaching may also alter the structural integrity of the materials, potentially affecting their long-term photocatalytic behavior.
The XRD patterns of the 6-CuH and 5-CuH500 samples after four consecutive uses, compared with their respective pristine forms, are presented in Fig. 23A. As shown, no significant changes were observed in the characteristic hydrotalcite-like layered structure or the oxide phases of either material after repeated use. These results indicate that both 6-CuH and 5-CuH500 retain their layered double hydroxide-like structural features and oxide phase compositions even after four reuse cycles.
Moreover, the concentration of Cu2+ ions leached into the solution increased progressively with each reuse cycle. In general, the total Cu2+ concentration detected after 180 minutes of irradiation was consistently higher for the 6-CuH sample (Fig. 23B) compared to the 5-CuH500 sample (Fig. 23C). This suggests that the Cu(OH)2 phase present in 6-CuH is more prone to dissolution than the CuO phase in 5-CuH500, despite the higher solution pH used for 6-CuH (pH = 4.15) compared to 5-CuH500 (pH = 3.0). Nevertheless, the photocatalytic performance of both materials remained relatively stable after three reuse cycles, achieving treatment efficiencies of 93.7% and 81.2% for 6-CuH and 5-CuH500, respectively, after 180 minutes of light exposure.
In general, Cu2+-modified Mg–Al LDHs possess considerable environmental benefits, including high treatment efficiency, and reusability. Nevertheless, material stability, prevention of metal leaching, and cost assessment remain essential to practical application.
(ii) Recommendations for future study future studies should optimize the Cu/Al ratio and investigate co-doping to prevent pore collapse. Enhancing structural stability during calcination and under extreme pH is critical to preserving material integrity. Minimizing Cu leaching through surface passivation is essential to reduce secondary pollution. Finally, assessing scalability, cost-efficiency, and reusability under real wastewater conditions will support practical application.
(iii) Environmental and industrial relevance these materials provide an efficient, sustainable solution for textile wastewater treatment under mild conditions (30 W LED, ambient temperature). They combine high adsorption with visible-light photocatalysis, reducing dependence on UV systems. To enable large-scale use, Cu leaching control and scalable synthesis remain key challenges. Overall, Cu–Mg–Al hydrotalcites present a viable alternative to conventional adsorbents and catalysts in eco-friendly wastewater treatment.
Additional figures and tables supporting the adsorption experiments and material characterization results. See DOI: https://doi.org/10.1039/d5ra02645h.
This journal is © The Royal Society of Chemistry 2025 |