Open Access Article
Sadia Sharmin Sporsho†
a,
Dipu Saha†b,
Mahmudul Hasan Khan†
b,
Md Shahriar Rahman†b,
Mahe Rukhb,
Md Reazul Islamb,
Tulie Chakmab,
Faysal Haquec,
Hridoy Royb,
Dipayan Sarkar
b and
Md Shahinoor Islam
*bd
aDepartment of Pharmaceutical Sciences, North South University, Dhaka 1229, Bangladesh
bDepartment of Chemical Engineering, Bangladesh University of Engineering and Technology (BUET), Dhaka-1000, Bangladesh. E-mail: shahinoorislam@che.buet.ac.bd
cDepartment of Mechanical Engineering, Bangladesh University of Engineering and Technology (BUET), Dhaka-1000, Bangladesh
dDepartment of Textile Engineering, Daffodil International University, Dhaka 1341, Bangladesh
First published on 16th December 2025
Pharmaceutical compounds have emerged as surface and groundwater contaminants over the last three decades. It is paramount to efficiently remove these contaminants from wastewater, as these molecules pose a severe threat to biodiversity and human health due to the inefficacy of wastewater treatment plants in removing many of these compounds, bioaccumulation in animal tissues, and harmful health effects caused at low concentrations. Although different removal techniques can be effective depending on the target compounds and wastewater characteristics, adsorption has a slight edge due to its low adsorbent and operational costs, high efficacy, and minimal byproducts. However, over the last decade, only a few articles have comprehensively reviewed the removal of pharmaceutical compounds through adsorption. This paper focuses on the environmental impact, detection accuracy, and effectiveness of various adsorbents for different pharmaceutical compounds. It critically analyzes the adsorption isotherms, adsorption kinetics, adsorption thermodynamics, and mechanisms of different adsorbents. Pore filling, electrostatic attraction, hydrophobic interactions, surface complexation (or bond formation), hydrogen bonding, and π–π interactions are the primary mechanisms for target molecule removal during adsorption. The sustainability metrics of different adsorbents are explored for scale-up, as well as effective strategies for managing used adsorbents to support sustainability, covering the gap from the lab scale to the industrial scale.
A wide variety of water sources have been found to contain almost every category of drugs and pharmaceuticals. Antibiotics, β-blockers, steroids, analgesics, anti-diabetics, anti-depressants, anti-epileptics, antihistamines, anti-psychotics, cytostatic, gastrointestinal, and lipid regulators are among the various types of medications and pharmaceuticals found in water.12,13 Antibiotics receive special attention within the pharmaceutical industry due to their role in promoting the emergence and dissemination of antibiotic resistance among microorganisms, particularly pathogens, through environmental contamination. The concentrations of antiviral drugs and antibiotics (e.g., azithromycin) in wastewater sharply increased (>70%) during the pandemic.14–16 Most of the pharmaceutical components resist contemporary wastewater treatment methods and have slow biodegradation.17–20 It is also alarming that half of pharmaceutical wastewater is released into the environment without treatment.21 Thus, the pollutants make their way to the domestic water body. A few of them can withstand water for an extended period. In most cases, the levels of residues from newly emerging contaminants are measured in micrograms per liter.
Pollutants in wastewater can be removed by combining physical, chemical, and biological treatment methods. Chemical treatments include coagulation, chemical oxidation, advanced oxidation, and electrochemical treatment.22 Advanced oxidation processes (AOPs) are suitable for removing chloroquine,23 ivermectin,24 azithromycin,24 penicillin,25 diclofenac,26 ciprofloxacin, and paracetamol.27 However, these removal processes are primarily laboratory-based and costly. Chemical oxidation and electrochemical processes may form byproducts that might be more harmful and toxic than the primary compounds in wastewater. Biological treatments are ineffective and slow processes, as antibiotics are difficult for microorganisms to degrade. The physical treatment methods incorporate sedimentation, sand filtration, adsorption, and membrane treatments.28 Physical wastewater treatment facilities, such as sedimentation and sand filtration, cannot fully degrade pharmaceuticals due to their design, which typically handles organics in the mg L−1 range. Membrane treatments are highly effective in removing pharmaceutical compounds; however, cost, clogging, and the need for frequent cleaning are the major issues associated with these treatments. Nano-filtration can remove up to 85% of anti-inflammatory drugs from wastewater.28 Several downsides are associated with most systems, including time consumption, rigorous operating specifications, cost, and periodic maintenance.29–32 Considering these drawbacks, the adsorption technique is highly utilized due to its low cost, ease of operation, efficacy, and stability in removing pharmaceutical waste.30–33 Fig. 1 presents the number of articles published recently with keywords related to adsorption and pharmaceutical wastewater.
![]() | ||
| Fig. 1 Recently published articles on adsorption and pharmaceutical wastewater from 2020–2025 (language limited to English). Source: Scopus. | ||
Many kinds and categories of adsorbents have emerged from the macro-to the nanoscale in recent years. Fig. 2 presents a schematic overview of the principal approaches for treating pharmaceutical wastewater. The advantages of adsorbents are their small size, great potential for tuning according to needs, large surface area per unit mass, existence of multiple active sites, and high effectiveness in removing pollutants.34
The high binding capacity of adsorbents for pharmaceuticals has made adsorption a valuable method for purifying pharmaceutical effluents. Adsorption has shown potential for pharmaceutical removal from water and effluents to prevent pollution and waste handling because of its straightforward design, easy operation, and absence of unwanted byproducts.35 Furthermore, the cost of treating wastewater using adsorption can be comparable to other techniques, such as the advanced oxidation process, depending on the pollutants and employed technique.36–38 Recently, numerous studies have removed pharmaceuticals from wastewater using various adsorbents, including activated carbon, biochar, porous carbon, zeolite, MOF, graphene, polymer, perovskite, etc.35,39,40 Features of the pollutants, such as charge, shape, size, and solubility, greatly influence the binding of pollutant species to an adsorbent surface. Adsorbent-based pharmaceutical treatment has a facile design and requires less energy than the advanced oxidation process to remove pharmaceuticals.41–43 Advanced oxidation processes can produce toxic byproducts during operation, making it difficult to scale up.41,43 On the other hand, adsorbent-based treatment methods are easier to scale and modify for use in real-world applications. Adsorption-based pharmaceutical wastewater treatments are available in real-world applications. The most common companies are DESOTEC, HYERA INC., and NORIT, and they utilize activated carbon as an adsorbent, either in powdered or granular form, to remove pharmaceuticals.44–46
Several publications on degrading and removing pharmaceutical substances have recently appeared in top peer-reviewed journals. However, only a few published studies have detailed the use of adsorbents to remove these compounds. In this article, a comprehensive approach was taken to review the current state-of-the-art methods using next-generation adsorbents to remove emerging pharmaceutical contaminants, as well as future remedial methods available to achieve these treatments in a more eco-friendly and sustainable manner. The article also explores the adsorption isotherms, kinetics, and mechanisms of pharmaceutical waste removal.
A wide range of pH values, from 3.7 to 14, has been reported in various studies for different PWs (Table 1). Understanding the correlation between the pH and the pollutant can be crucial for efficient wastewater treatment. Like other pollutants, pharmaceutical pollutants (PP) are present in their ionic form in wastewater. The zeta potential (ZP) value indicates the electrostatic interaction between a charged surface and PP. The variation of ZP of PP is influenced by the pH of the solution, which affects the PP removal kinetics and mechanism.51 Chemical oxygen demand (COD) and biochemical oxygen demand (BOD) are the main water quality parameters. The typical COD and BOD content in wastewater from pharmaceutical manufacturing industries ranges between 800 and 60
000 mg L−1 and 40 and 21
560 mg L−1, respectively (Table 1).
| Manufacturing type | Main API | Main API concentration (mg L−1) | COD (mg L−1) | BOD5 (mg L−1) | TOC (mg L−1) | NH3–N (mg L−1) | pH | TSS (mg L−1) | TDS (mg L−1) | NO3− (mg L−1) | PO43− (mg L−1) | Reference |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| a NO3–N.b Total phosphate.c PO4–P.d NH4+; NS: not studied. | ||||||||||||
| Formulation facility | Carbamazepine | 0.84 (±0.19) | 4765 (±1405) | 634 (±100) | 1698 (±308) | 23.5 (±8) | 10.2 (±0.9) | NS | NS | NS | NS | 56 |
| Venlafaxine | 11.72 (±2.2) | |||||||||||
| Manufacturing industry | Paracetamol | 48 | 3500 | 466 | NS | NS | 6.8 | 360 | 600 | NS | NS | 57 |
| Bulk manufacturing unit | NS | NS | 37 410 (±225) |
21 560 (±160) |
8250 (±145) | NS | 6780 (±180) | 21 340 (±450) |
18 (±1.25) | 0.2 (±0.03) | 58 | |
| Bulk manufacturing unit | Antibiotic | NS | 15 365 (±1214) |
NS | 7624 (±710) | NS | 7–8 | 388 (±87) | 22 168 (±3757) |
NS | NS | 59 |
| Chemical synthesis | Antibiotic | NS | 16 249 (±714) |
NS | 6697 (±1047) | NS | 7–8 | 199 (±59) | 29 450 (±1209) |
NS | 188 (±29) | 60 |
| Chemical synthesis | Etodolac | 50–215 | 20 000–23000 |
NS | NS | NS | 3.7–11.3 | NS | 30–34a | 0.5–2.2b | 61 | |
| Manufacturing and equipment cleaning | Penicillin | NS | 16 547 (±1827) |
10 184 (±2574) |
8083 (±578) | 72 (±46) | 7–9 | 285 (±175) | 24 899 (±1758) |
NS | NS | 47 |
| Triethylamine | 9872 (±2142) | |||||||||||
| Chemical synthesis | Antibiotic | NS | 39 000–60000 |
NS | NS | NS | 7–8 | 800–1000 | NS | NS | 3–6c | 62 |
| Hospital water treatment plant | NS | NS | 376 | NS | NS | 22.3d | 7–7.5 | NS | NS | 0.2 | NS | 63 |
| Chemical synthesis | Diclofenac | 18–20 | 500–593.53 | NS | 170.76–200 | NS | 7–7.14 | NS | NS | NS | NS | 64 |
| Chlorophenol | <3 | |||||||||||
| Wastewater treatment plant | NS | NS | 1800 | 1130 | NS | NS | 7.5 | 750 | NS | NS | 16 | 52 |
| Pharmaceutical industry | NS | NS | 810 | 40 | NS | NS | 7.27 | 118 | NS | 7.23 | 2.13 | 65 |
| Hospital wastewater | NS | NS | 918 | 400 | 220 | 46 | 7.7 | 310 | NS | NS | 7.4 | 65 |
| Fermentation-based PW | NS | NS | 6800.5 | >2040 | 2476.3 | 50.7 | NS | 188.3 | NS | NS | NS | 66 |
The most frequently detected active pharmaceutical ingredients (APIs) in wastewater are blood lipid regulators, non-steroidal anti-inflammatory drugs (NSAIDs), antibiotics, selective serotonin reuptake inhibitors (SSRIs), analgesics, β-blockers, hormones, and antihistamines.52 Kostich et al.53 investigated the effluents of the 50 largest wastewater treatment plants in the US and reported 56 APIs present in various concentrations. Among these APIs, the most detected ones (30 detections) are atorvastatin, carbamazepine, ciprofloxacin, diltiazem, furosemide, diltiazem-desmethyl, gemfibrozil, hydrochlorothiazide, metoprolol, ofloxacin, oxycodone, propranolol, sertraline, sulfamethoxazole, triamterene, trimethoprim, valsartan, and verapamil. Valsartan was found to have the highest maximum concentration of 5300 ng L−1, followed by ibuprofen (4200 ng L−1), lisinopril (3300 ng L−1), atenolol (3000 ng L−1), sulfamethoxazole (2900 ng L−1), hydrochlorothiazide (2800 ng L−1), and gemfibrozil (2300 ng L−1). The USGS surveyed 1091 sites spanning 46 states to assess the pharmaceutical and hormone content of groundwater from 2013 to 2015.54 1,7-dimethylxanthine, carbamazepine, meprobamate, sulfamethoxazole, and bisphenol A showed the highest number of detections (more than 0.5%) with detection counts of 9, 18, 8, 12, and 7, respectively. This study revealed that principal aquifers in the USA have already been contaminated with pharmaceutical and hormone compounds, although at levels below safety benchmarks for humans. Another USGS survey was conducted (2014–2017) on 308 wadable streams across four regions of the USA to measure 108 pharmaceutical analytes.55 Compounds detected in the most significant number of sites were nicotine (70% sites), metformin (68% sites), cotinine (47% sites), lidocaine (42% sites), caffeine (42% sites), carbamazepine (41% sites), and acetaminophen (26% sites).
The various active ingredients of pharmaceutical pollutants have a significant impact on aquatic organisms, posing a severe threat to their aquatic life. The ecotoxicological impact of pharmaceutical compounds on different organisms is summarized in Table 2. The literature review in this section highlights the severe threat different pollutants pose to aquatic creatures. Although most of the work reviewed here does not represent environmentally relevant conditions, the bioaccumulation capabilities of these pollutants can potentially create similar complications under low pollutant concentrations. Therefore, we need to understand various wastewater treatment techniques and implement effective methods to minimize potential environmental damage.
| Active ingredient | Structure | Classification | Organism | Lethal concentration, 50% (LC50) (mg L−1) | Effective concentration, 50% (EC50) (mg L−1) | Effect | Reference |
|---|---|---|---|---|---|---|---|
| Fluoxetine | ![]() |
Selective serotonin reuptake inhibitor (SSRI) | Oryzias javanicus | 1.23 (96 h) | 0.01 (96 h, swimming behavior) | Decreased swimming performance, disruption of the endocrine system, DNA damage, and oxidative and endoplasmic reticulum stress | 69 |
| Hydra magnipapillata | 2.90 (72 h) | Reduced reproduction, morphological deformation, neurotransmission disturbance, DNA damage, and oxidative stress | |||||
| Sertraline | ![]() |
SRRI | Daphnia magna | 0.12 (21 days) | 0.06 (21 days, reproduction) | Increased mortality and reduced reproduction | 70 |
| Carbamazepine | ![]() |
Anticonvulsant | Oncorhynchus mykiss | Two-fold reduction of intestinal RNA/DNA ratio, induced oxidative stress, and disrupted osmoregulation | 71 | ||
| Gemfibrozil | ![]() |
Blood lipid regulator | Danio rerio larvae | 11.01 (96 h) | Increased hatching time, reduced hatchability, locomotion impairment, edema formation, and yolk sac malformation | 72 | |
| Diclofenac | ![]() |
Non-steroidal anti-inflammatory drug (NSAID) | Mytilus galloprovincialis | DNA damage, oxidative stress, lipid metabolism disorder, and osmoregulation disruption | 73 | ||
| Ibuprofen | ![]() |
NSAID | Planorbis carinatus | 17.1 (72 h) | Decrease in growth | 74 | |
| Dreissena polymorpha | Induction of oxidative stress and exhibited genotoxic effects | 57 | |||||
| Aspirin | ![]() |
NSAID | Daphnia | 86.1 (48 h, daphnid immobilization) | DNA damage, deformities in neonates, behavioral and physiological changes, induction of oxidative stress, and reduced reproduction | 75 | |
| Erythromycin | ![]() |
Antibiotic | Oncorhynchus mykiss | Gill: vasodilation, oedema, epithelial lifting, lamellar epithelial desquamation, lamellar fusion, hyperplasia on the epithelium of the gill | 76 | ||
| Liver: increase in sinusoidal space, hemorrhage, cytoplasmic vacuolization, nuclear degeneration, and nuclear/cellular hypertrophy of hepatocytes |
Pharmaceuticals, including antibiotics, hormones, and analgesics, are increasingly being detected in water bodies, disrupting aquatic ecosystems and contributing to issues like antimicrobial resistance (AMR).67 An adsorption system has been successfully demonstrated for removing pharmaceuticals from real wastewater using low-cost sorbents in a pilot-scale plant.68 The cost-effective nature and exceptional performance of the adsorption technique provide a distinct advantage over traditional methods. Therefore, the primary objective of this review is to investigate the role of adsorbents in removing pharmaceuticals from aquatic systems.
| Adsorption kinetics model | Linear form | Non-linear form | Application insights | Adsorption isotherm model | Linear form | Non-linear form | Application insights |
|---|---|---|---|---|---|---|---|
| a Note: t is the adsorption time (min), qt (mg g−1) is the adsorbed amount of the adsorbate at time t, qe is the adsorption capacity at equilibrium (mg g−1), k1 is the pseudo-first-order rate constant (min−1), k2 is the pseudo-second-order rate constant (g mg−1 min−1), ki (mg g−1 min−1/2) is the rate constant of the intraparticle diffusion model, c is the degree of diffusion, α (mg g−1 min−1) is the primary rate of adsorption, β represents the desorption parameter in the Elovich kinetic model, Bt is a parameter relating to adsorbent characteristics, F(t) is defined as qt/qe, KF (L1/n mg1−1/n g−1) is the Freundlich constant, n is the dimensionless Freundlich intensity parameter, KDR is the D–R constant related to the mean free energy of adsorption, ε is the adsorption potential, KRP is the Redlich–Peterson constant, which is related to the adsorption capacity, αRP is the Redlich–Peterson constant, related to the adsorption intensity, β is an exponent that lies between 0 and 1, A is the Temkin isotherm equilibrium binding constant (L g−1), which is related to the maximum binding energy, qms (mg g−1) is the maximum adsorbed amount, Ks (Lns mg−ns), and ns is the Sips constant. | |||||||
| Pseudo-first-order model (Lagergren model) | ln(qe − qt) = lnqe − k1t | qt = qe(1 − e−k1t) | Describes initial rapid adsorption; suitable for low-concentration or early-time studies | Langmuir model | Ce/qe = q/(KL qL)+ Ce/qL | ![]() |
Estimates maximum adsorption capacity (qmax); guides adsorbent dosage and saturation prediction |
| Pseudo-second-order model | t/qt = 1/(k2qe2) + t/qe | ![]() |
Predicts overall adsorption capacity and equilibrium; widely used for pharmaceutical uptake | Freundlich model | ln qe = ln KF + (1/n)ln Ce |
![]() |
Evaluates surface heterogeneity; predicts adsorption intensity in multi-contaminant wastewater |
| Intraparticle diffusion model (Weber–Morris model) | qt = kit1/2 + c | — | Identifies pore-diffusion control and multi-stage adsorption, important for porous adsorbents | Dubinin–Radushkevich (D–R) model | ln qe = ln qD − KDRε2 |
![]() |
Distinguishes physisorption (<8 kJ mol−1) and chemisorption; indicates regeneration feasibility |
![]() |
|||||||
| Pore-diffusion model | log qt = k log t |
— | Distinguishes film-controlled vs. diffusion-controlled processes | Redlich–Peterson model | ![]() |
![]() |
Provides accurate fits for real systems; useful for industrial-scale predictions |
| Elovich model | ![]() |
— | Explains non-uniform surface energy adsorption; useful for biochar and metal-doped adsorbents | Temkin model | ![]() |
![]() |
Indicates chemisorption on heterogeneous surfaces |
| Boyd's film diffusion model | ![]() |
![]() |
Distinguishes film-controlled vs. particle-controlled adsorption, key for column scale-up | Sips model | ![]() |
![]() |
Ideal for heterogeneous, multi-site adsorbents; effective across low and high concentration ranges |
The adsorption performance in fixed-bed systems is typically evaluated using dynamic modeling, which predicts breakthrough curves and service life under realistic flow conditions. Among the widely used models, the Thomas, Bohart–Adams, and Yoon–Nelson equations are the most common due to their simplicity, robust applicability, and ability to guide industrial-scale column design.
Thomas model: The Thomas model assumes plug-flow behavior with second-order reversible kinetics and negligible axial dispersion. It is widely applied to predict adsorption capacity, breakthrough curves, and bed exhaustion times. The model is given by:
Bohart–Adams model: The Bohart–Adams model correlates column performance with bed depth and service time, assuming that adsorption is primarily controlled by surface reaction kinetics. It is beneficial for estimating bed depth requirements, service time, and early breakthrough performance. The linearized form is:
Yoon–Nelson model: The Yoon–Nelson model simplifies column design by assuming that the adsorbate breakthrough probability is directly proportional to the adsorption rate. This model predicts the time for 50% breakthrough (τ) without requiring extensive parameter fitting:
Chemisorption, in contrast, occurs when the adsorbate and adsorbent share electrons, forming strong valence bonds. This adsorption type is typically irreversible, highly selective, and characterized by significant thermal effects. Chemisorption is characterized by elevated adsorption energies, typically ranging from 80 to 400 kJ mol−1, and often results in structural changes in the adsorbate or the formation of stable surface complexes. Principal chemisorption mechanisms encompass ion-pair and electron–transfer interactions between charged pharmaceutical entities and active surface sites, surface complexation and coordinate bonding at metal centers, ion exchange, and, in certain instances, redox reactions that augment binding or initiate partial transformation of the pollutant.87,88 Robust and selective interactions are essential for the elimination of persistent and low-concentration medicines from intricate aqueous matrices. Analyzing the adsorption mechanism provides substantial insight into the efficacy of various adsorbents for different pharmaceutical adsorbates. This review thoroughly explores the interactions between pharmaceutical pollutants and adsorbents during the adsorption process. The adsorption of pharmaceutical micro-contaminants onto an adsorbent surface primarily occurs due to surface energy, as the atoms or functional groups on the adsorbent attract the adsorbate to minimize surface energy. The adsorption mechanism is significantly influenced by ambient and material variables, including pH, ionic strength, surface charge, and the type and density of surface functional groups. Functional groups like –OH, –NH2, –COOH, and –C
O promote hydrogen bonding, ion exchange, and coordination with medicinal compounds, while the hydrophilicity or hydrophobicity of the adsorbent surface determines its affinity for polar or nonpolar pollutants. The pH of the solution regulates the ionization of both adsorbates and adsorbent surfaces; therefore, it directly influences electrostatic and ion-pair interactions. The existence of concurrent ions or natural organic materials can also influence adsorption mechanisms by competing for active sites or obstructing electrostatic interactions. The driving force behind adsorption results from the combined effect of multiple interactions that contribute to the total free energy of the process. These interactions include hydrogen bonding, electrostatic attraction, π–π interactions, and dipole–dipole interactions between the adsorbent and adsorbate.89 In certain instances, van der Waals forces and hydrophobic interactions contribute to the adsorption of organic molecules onto adsorbent materials. Van der Waals forces refer to intermolecular attractions, categorized as weak London dispersion forces and stronger dipole–dipole interactions. Another possible binding mechanism for pharmaceutical contaminants is the hydrophobic interaction between nonpolar groups. In contrast to intermolecular forces, hydrophobic interactions are driven by entropy, resulting from the exclusion of chemicals from the aqueous phase rather than a direct attraction to the adsorbent. The potential adsorption mechanisms of aqueous pharmaceuticals onto various adsorbents are summarized in Fig. 3.
![]() | ||
| Fig. 3 Potential mechanisms behind the adsorption of pharmaceuticals onto nanoparticles (adapted from ref. 90 with permission from Springer Nature, copyright 2023). | ||
Izzo et al. successfully modified natural zeolites using long-chain cationic surfactants to develop a composite material with a high adsorption capacity for ibuprofen.92 Martucci et al. investigated the adsorption capacity for the removal of erythromycin, carbamazepine, and levofloxacin drugs from pharmaceutical wastewater using three organophilic zeolites (Y, mordenite (MOR), ZSM-5).93 Zeolite Y (dealuminated faujasite) was studied by Braschi et al. and confirmed to effectively remove sulfonamide antibiotics from water, which significantly contribute to bacterial resistance.94 The adsorptive removal of norfloxacin (NOR) and ofloxacin (OFL) was explored by Zhao et al. using a polyethylene glycol (PEG-4000) surfactant-modified and zeolite-supported nanoscale zero-valent iron composite.95 Fig. 4 shows the various steps involved in the synthesis, characterization, and removal mechanism of two antibiotics with PZ-NZVI composite. Within one hour, more than 95% of NOR or OFL could be removed from the solution using PZ-NZVI, and the adsorption process could be best described using the pseudo-second-order kinetic model and the Temkin isotherm model. The characterization results before and after adsorption, as well as batch studies, demonstrated that various processes, including hydrophobic interaction, bidentate complex formation between Fe and fluoroquinolones, pore filling, and electrostatic interaction, can control the sorption process.
![]() | ||
| Fig. 4 A graphical representation of removing two fluoroquinolone antibiotics (NOR: naproxen and OFL: ofloxacin) using PEG-4000-stabilized nanoscale zero-valent iron supported on zeolite (PZ-NZVI) (adapted from ref. 95 with permission from Elsevier, copyright 2020). | ||
Arabkhani et al. reported an ultra-high adsorption capacity value of 2594 mg g−1 at 30 °C for the removal of diclofenac sodium from pharmaceutical wastewater by synthesizing graphene oxide (GO) nanosheets with zeolitic imidazolate framework-8 (ZIF-8), pseudo-boehmite (γ-AlOOH), and iron oxide (Fe3O4) nanoparticles.96 Liu et al. confirmed that modified zeolite-supported nano-MoS2 (MoS2@zeolite) with multiple adsorption sites is an efficient and promising adsorbent for treating pharmaceutical wastewater tetracycline.97 Attia et al. synthesized magnetic nanoparticles coated with zeolite (MNCZ) to adsorb medicinal substances from pharmaceutical compounds.98 Hexadecyltrimethylammonium (HDTMA)-modified zeolites showed hydrophobic interaction mechanisms when estrogenic contaminants were removed from wastewater.99 Thus, adsorbents based on zeolites and nanocomposites might be promising next-generation adsorbent materials for treating pharmaceutical wastewater; however, continuous operation will not eliminate the need for regular adsorbent material replacement and regeneration.
MIL-101 (chromium–benzene dicarboxylate), in which MIL stands for Material of Institute Lavoisier, is one of several MOFs created so far that have been extensively studied for prospective use for the removal of naproxen and clofibric acid from wastewater due to its very high porosity (1.9 cm3 g−1). Hasan et al. further functionalized MIL-101 with an acidic group (AMSA-MIL-101) and a primary group (ED-MIL-101). They conducted batch experiments to explore the adsorption effectiveness of eliminating naproxen and clofibric acid.100 UiO-66 with controlled defects contained more open frameworks and showed a higher affinity for diclofenac than other pharmaceuticals.101 Methanol-activated Cu-based MOF(HKUST-1) showed excellent adsorption capacity to remove sulfonamide antibiotics and sulfachloropyridazine (SCP).102 The high surface area, large pore volume, and unsaturated metal sites resulted in faster, spontaneous, and endothermic adsorption kinetics for removing sulfonamide antibiotics. Fig. 5 shows that electrostatic interactions, H-bonding interaction with the H of the NH2 from the SCP and the oxygen of the HKUST-1 clusters, and π–π stacking between the benzene ring of the MOF and the SCP are primarily responsible for the high adsorption capacity.
![]() | ||
| Fig. 5 Adsorption mechanism of SCP on HKUST-1, (a) H-bond formed with H of NH2 from SCP and oxygen of HKUST-1 clusters, (b) H-bond between Cu–O–Cu and H of amide group is indicating the removal of SCP in its neutral form (adapted from ref. 103 with permission from Elsevier, copyright 2016). | ||
Lu et al. fabricated a novel amino-functionalized aluminum-based metal–organic framework (Al-MOF@NH2), demonstrating high hydrocortisone (HC) adsorption capacity, a common steroid hormone drug. Large-scale manufacturing, water stability, and reusability are three critical obstacles to using MOFs as adsorbents. Most MOFs are not water stable, which may lead to poor recovery and even second-hand pollution from metal leaching.
Recent studies have reported that many MOFs, especially those with weak metal–ligand bonds, suffer from poor aqueous and thermal stability, which restricts their long-term use in water treatment.104 Additionally, metal ion leaching from MOFs—particularly those based on transition metals such as Cu, Zn, and Fe—poses risks to environmental safety and downstream processes.105 The potential toxicity of uncoordinated or degraded organic linkers, such as imidazolates and terephthalates, is also an emerging concern, especially under variable pH and oxidizing conditions.106 We have incorporated these findings into the discussion to present a more nuanced and realistic evaluation of MOFs' applicability in pharmaceutical wastewater treatment.
Graphene is a two-dimensional arrangement of carbon atoms organized in a hexagonal lattice structure, constituting a singular layer.107 Fig. 6 shows the two-dimensional (2D) structure resulting from sp2 hybridization of its carbon atoms arranged in a honeycomb framework.108 The excellent dispersion properties of graphene are due to the weak van der Waals forces that bind the layers (bond length 0.142 nm) together.
![]() | ||
| Fig. 6 (a) The sp2 hybridization arrangement of carbon atoms, which are closely packed together in a honeycomb lattice formation. (b) The atomic structure of graphene, emphasizing the individual carbon atoms and their connections within the lattice (adapted from ref. 109 with permission from Elsevier, copyright 2020). | ||
The increasing interest in graphene originates from its remarkable physicochemical attributes, including its elevated specific surface area (a theoretical surface area of 2630 m2 g−1),110 exceptional electrical and thermal conductivity,111 chemical structure, and mechanical strength. Graphene acts as a rapid adsorbent for diverse contaminants thanks to its extensive, delocalized π-electron system, enabling robust interactions with other pollutants. Saravanan et al. provided an in-depth analysis of the applications of materials derived from graphene in wastewater treatment as adsorbents, electrodes, and photocatalysts to efficiently remove harmful pharmaceutical pollutants, heavy metals, dyes, and other contaminants, as shown in Fig. 7.112
![]() | ||
| Fig. 7 A schematic representation of possible mechanisms for water pollutant removal by graphene adsorbents, photocatalysts, and electrodes (adapted from ref. 112 with permission from Elsevier, copyright 2022). | ||
The application of graphene-based materials as photocatalysts for removing decomposing organic contaminants from pharmaceutical effluent has been reported.113 Graphene can enable the production of reactive oxygen species when paired with appropriate photocatalytic substances, such as metal oxides or semiconductors, and subjected to light irradiation.114 These reactive species can facilitate the degradation of organic molecules into less detrimental chemicals via oxidation mechanisms. Jauris et al. investigated the adsorption behavior of sodium diclofenac (s-DCF) on several carbon-based materials, including pristine graphene, graphene with a vacancy, reduced graphene oxide (rGO), and functionalized graphene nanoribbons.115 The primary objective of this research was to gain insights into the underlying process of s-DCF adsorption on the carbon lattice. The computer simulations demonstrated that the interactions between pristine graphene and s-DCF can be attributed to a physical adsorption mechanism. However, in the case of pristine graphene and graphene with a single vacancy, the outcomes indicated the presence of π–π interactions.
Iron-based nanoparticles have been extensively studied in various forms, including doped, composite, and spinel oxides. Various chemical methods have been reported for synthesizing pristine/doped/composite iron oxide nanoparticles, including co-precipitation, sol–gel, thermal decomposition, hydrothermal, and polyol methods. Other physical methods include solid-state ball milling, gas phase deposition, and pulsed-laser ablation. Hematite (α-Fe2O3), among various iron oxide polymorphs, has garnered significant attention due to its exceptional resistance to corrosion, non-toxic nature, high stability in atmospheric conditions, and environmentally friendly properties. The particle size, shape, and composition of chemically synthesized iron nanoparticles (NPs) are influenced by various factors, including the precursor salt type, Fe(II) to Fe(III) ratio, pH, and ionic strength. In a study by Ali et al., five types of adsorptive removal mechanisms were outlined for iron-based nanoparticles.117 These mechanisms include the electrostatic interaction between pollutants and magnetic nanomaterials, facilitated by diverse biomolecules present on the surfaces of magnetic nanoparticles. Chemical diffusion occurs between the adsorbent and adsorbate, while surface precipitation, redox reactions, and ion exchange are also significant mechanisms. Hydroxyl (OH) functional groups play a crucial role in the ion exchange process. The tendency of different groups of antibiotics to dissociate into cations, zwitterions, and anions at varying pH levels cannot make their mechanism of adsorption a stereotype. Hence, in addition to experimental techniques, density functional theory calculations can provide crucial insights into the mechanism of removing contaminants.
| Adsorbent | Synthesis techniques | Characterization techniques | Adsorbate | Surface area (m2 g−1) | Experimental conditions | Maximum experimental adsorption capacity (mg g−1) | Adsorption capacity per unit surface area (mg m−2) | Adsorption isotherm | Adsorption kinetics | Thermodynamic parameters (ΔG kJ mol−1, ΔH kJ mol−1, and ΔS J mol−1 K−1) | Key insights | Reference |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Zeolite (ZPC800) | Microwave-assisted solvothermal method, pyrolysis temperature 800 °C | Surface: BET | Tetracycline | 268 | pH 6, initial adsorbent dosage = 100 mg L−1, contact time = 300 min | 317.50 | 1.18 | Temkin | Elovich | — | Heterogeneous surface adsorption, non-uniform surface energy adsorption | 150 |
| Microwave-assisted solvothermal method, pyrolysis temperature 800 °C | Structural morphology: SEM, adsorption quantification: UV-vis | Ciprofloxacin | pH 8, initial adsorbent dosage = 40 mg L−1, contact time = 350 min | 270.67 | 0.99 | Temkin | Pseudo-second-order | — | Chemisorption on heterogeneous surface | |||
| Others: XRD, FTIR | ||||||||||||
| Zeolite (HSZ-690HOA zeolite) | Commercial high silica zeolite mordenite, calcination temperature 800 °C | Surface: BET | Sulfachloropyridazine | 453 | At 25 °C and 65 °C, 4 h contact time | 151 | 0.33 | N/A | N/A | — | Hydrogen bonding | 151 |
| Adsorption quantification: HPLC-UV, HPLC-MS | ||||||||||||
| Others: NMR, TGA | ||||||||||||
| Zeolite (faujasite zeolite Y) | “Y” type faujausite zeolite powder with a 200 SiO2/Al2O3 (mol mol−1) ratio | Surface: BET, others: TGA, XRD | Sulfonamide | 852 | At 25 °C and 65 °C, pH 5.8, 24 h contact time | 280 | 0.33 | N/A | N/A | — | N/A | 94 |
| Zeolite (ZIF-8) | Sonochemical synthesis, activated by heating under vacuum at 80 °C | Surface: BET | Ofloxacin | At pH 7.7, 10 mg adsorbent L−1 | 95 ± 10 | 0.31 | Langmuir | Pseudo-second-order | — | Chemisorption | 152 | |
| Structural morphology: SEM | ||||||||||||
| Adsorption quantification: HPLC-UV, HPLC-FD | ||||||||||||
| Others: XRD, FTIR | ||||||||||||
| Zeolite (zeolite Y) | Sol–gel method, calcination temperature 500 °C | Surface: BET | Tetracycline | 657.44 | At 30 °C, pH 6.7 | 201.77 | Langmuir | Pseudo-second-order | — | Chemisorption | 153 | |
| Structural morphology: SEM | ||||||||||||
| Others: XRF, XRD, FTIR | ||||||||||||
| Zeolite (PZ-NZVI) | One-pot method | Surface: BET | Norfloxacin, ofloxacin | 26.48 | At pH 4–10 | 54.67 mg g−1 for norfloxacin and 48.88 mg g−1 for ofloxacin | 2.06 mg m−2 for norfloxacin and 1.85 mg m−2 for ofloxacin | Langmuir | Pseudo-second-order | −27.2, −6.98, 70.3 for NOR -26.63, −17.97, 30.3 for OFL | Electrostatic, hydrophobic interaction and complexation, spontaneous and exothermic adsorption | 95 |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: XPS, XRD, FTIR, TGA | ||||||||||||
| Zeolite (magnetic GO/ZIF-8/AlOOH-NC) | Combining the solvothermal and solid-state dispersion method | Surface: BET | Diclofenac sodium | 777 | At 30 °C, pH 7.5, equilibrium time 50 min | 2594 | 3.33 | Langmuir | Pseudo-second-order | — | H-bonding, electrostatic attraction, n-pi interaction, π–π interaction, physical interaction | 96 |
| Structural morphology: FE-SEM, EDS | ||||||||||||
| Others: AFM, XRD, FTIR, TGA | ||||||||||||
| Zeolite (CTAB-ZIF-67) | — | Surface: BET | Diclofenac sodium | 709–1103 | At pH 5–10, 90 min equilibrium time | 60.60 | 0.08 | Langmuir | Pseudo-second-order | −29.53, 40, 240 | Electrostatic attraction, spontaneous, and endothermic | 154 |
| Structural morphology: FE-SEM, EDS | ||||||||||||
| Others: Zeta potential, XRD, FTIR, TGA | ||||||||||||
| Zeolite (ZCPC) | Raw clinoptilolite-rich zeolitic tuff from Zlatokop deposit | Surface: BET | Diclofenac sodium | At pH 7.4, 60 min equilibrium time | 160 | Langmuir | N/A | — | N/A | 155 | ||
| Others: Zeta potential, FTIR, TGA | ||||||||||||
| Zeolite (MoS2@zeolite-5) | Combining the ultrasonic and hydrothermal method | Surface: BET | Tetracycline | 15.82 | At 35 °C, at pH 4 | 396.70 | 25.08 | Langmuir | Pseudo-second-order | −0.806, 66.17, 224.98 | Chelation, electrostatic action, π–π action, and H-bonding | 97 |
| Structural morphology: SEM, EDS, HR-TEM | ||||||||||||
| Others: XPS, XRD, FTIR | ||||||||||||
| Zeolite (MNCZ) | — | Surface: BET | Diclofenac-Na, naproxen, gemfibrozil and ibuprofen | At pH 2 | Removal efficiency > 99% | Freundlich | Pseudo-second-order | — | Chemisorption | 98 | ||
| Adsorption quantification: LC-MS/MS | ||||||||||||
| Zeolite (modified zeolite) | Modified by equilibrating RZ and HDTMA solutions | Surface: BET, adsorption quantification: HPLC | Estrone and 17β-estradiol | At pH 5–7, low temperature | 8.29 mg g−1 for estrone and 7.09 mg g−1 17β-estradiol | Langmuir | Pseudo-second-order | — | Distribution effects and surface adsorption | 99 | ||
| Structural morphology: SEM | ||||||||||||
| Others: Contact angle, FTIR, zeta potential | ||||||||||||
| Zeolite (FeO-NCP) | Dry milling method | Surface: BET | Tetracycline | At pH 5 | >50% degradation | N/A | N/A | — | N/A | 156 | ||
| Adsorption quantification: UV-vis | ||||||||||||
| Structural morphology: SEM, TEM | ||||||||||||
| Others: AAS, FTIR, XRD, TOC | ||||||||||||
| Zeolite (TiO2/zeolite) | Acid-activated and acid-alkali-activated zeolites | Surface: BET | Amoxicillin | 240 min irradiation time | >99% removal efficiency | N/A | N/A | — | N/A | 157 | ||
| Adsorption quantification: HPLC, LC-MS | ||||||||||||
| Structural morphology: SEM, EDS | ||||||||||||
| Others: XRD | ||||||||||||
| MOF (amino-functionalized MIL-88B(Fe)-based porous carbon) | Microwave-assisted solvothermal method | Surface: BET | Ciprofloxacin | 215.1 | At 700 °C pyrolysis, pH 6, adsorbent 0.1 g L−1, contact time 240 min | 102.5 | 0.48 | Freundlich | Pseudo-second-order | — | Chemisorption | 158 |
| Structural morphology: SEM | ||||||||||||
| Others: XRD, XPS | ||||||||||||
| MOF (MPC800) | Solvothermal method, pyrolysis temperature 800 °C | Surface: BET | Ciprofloxacin | 199 | At 800 °C pyrolysis, pH 4, adsorbent dosage 0.1 g L−1, ionic strength 0.4 mol L−1 | 90.9 | 0.46 | Langmuir | Pseudo-second-order | — | Chemisorption, electrostatic attraction | 159 |
| Structural morphology: SEM, TEM | ||||||||||||
| Others: XRD, FTIR, RAMAN | ||||||||||||
| MOF (Cr-based MOF (MIL-101)) | Hydrothermal method | Surface: BET | Naproxen, clofibric acid | 3014 | At 25 °C, pH 4.5, contact time 12 h, 100 g adsorbent L−1 | 132 mg g−1 for naproxen, 244 mg g−1 for clofibric acid | 0.04 mg m−2 for naproxen, 0.08 mg m−2 for clofibric acid | Langmuir | Pseudo-second-order | — | Electrostatic interaction | 100 |
| Others: XRD | ||||||||||||
| MOF (acidic Cr-based MOF (AMSA-MIL-101)) | Hydrothermal method | Surface: BET | Naproxen, clofibric acid | 2322 | At 25 °C, pH 4.5, 100 g adsorbent L−1 | 93 mg g−1 for naproxen, 105 mg g−1 for clofibric acid | 0.04 mg m−2 for naproxen, 0.05 mg m−2 for clofibric acid | Langmuir | Pseudo-second-order | — | Acid-base interaction | 160 |
| Others: XRD, FTIR, elemental analyzer | ||||||||||||
| MOF (basic Cr-based MOF (ED-MIL-101)) | Hydrothermal method | Surface: BET | Naproxen, clofibric acid | 2555 | At 25 °C, pH 4.5, 100 g adsorbent L−1 | 154 mg g−1 for naproxen, 347 mg g−1 for clofibric acid | 0.06 mg m−2 for naproxen, 0.14 mg m−2 for clofibric acid | Langmuir | Pseudo-second-order | — | Acid-base interaction | 160 |
| Others: XRD, FTIR, elemental analyzer | ||||||||||||
| MOF (al-based MOF[MIL-53(Al)]) | Hydrothermal method | Surface: BET | Nitroimidazole | 1401 | At 30 °C, pH 6.4, 1 g adsorbent L−1 | 467.3 | 0.33 | Langmuir | Pseudo-second-order | — | Van der Waals interaction | 161 |
| Structural morphology: STEM-HAADF, TEM | ||||||||||||
| Others: XRD, XPS, FTIR | ||||||||||||
| MOF (MIL-101-urea) | Hydrothermal method | Surface: BET | Nitroimidazole | 1970 | At 25 °C, 12 h contact time, pH 6.3 | 185 mg L−1 for DMZ, 188 mg g−1 for MNZ | 0.09 mg m−2 for DMZ, 0.1 mg m−2 for MNZ | Langmuir | N/A | — | H-bond | 162 |
| Others: XRD, elemental analyzer, FTIR | ||||||||||||
| MOF (methanol-activated Cu-based MOF (HKUST-1)) | Hydrothermal method, activated by methanol | Surface: BET | Sulfachloropyridazine | 1700 | At 25 °C, pH = 7.5, 0.1 g adsorbent L−1 | 384 | 0.22 | Langmuir | Pseudo-second-order | −28.8, 4, 110.3 | π–π stacking, H-bonding, electrostatic interaction, spontaneous, and endothermic adsorption | 102 |
| Structural morphology: SEM | ||||||||||||
| Others: XRD, TGA, FTIR | ||||||||||||
| MOF (chloroform-activated Zr-based MOF (UiO-66)) | Modified solvothermal method | Surface: BET | Sulfachloropyridazine | 1155 | At 25 °C, pH 5.5, 0.1 g adsorbent L−1 | 417 | 0.36 | Langmuir | Pseudo-second-order | −30.4, −60.6, −100.9 | Hydrophobicity, π–π interaction, electrostatic interaction, spontaneous, and endothermic adsorption | 163 |
| Structural morphology: SEM | ||||||||||||
| Others: XRD, TGA, FTIR | ||||||||||||
| MOF (CFC/UiO-66-NH2/AgI) | Microwave-assisted hydrothermal method | Surface: BET | Levofloxacin, ciprofloxacin | 730.8 | At 4.5–8.5 pH, 10 mg adsorbent L−1 | Degrade 84.5% levofloxacin, 79.6% ciprofloxacin | N/A | N/A | — | N/A | 164 | |
| Adsorption quantification: UV-vis, LC-MS | ||||||||||||
| Structural morphology: FE-SEM, TEM, HR-TEM | ||||||||||||
| Others: XRD, XPS | ||||||||||||
| MOF (copper meso-tetra(4-carboxyphenyl) porphine-MOFs) | Modified solvothermal method | Surface: BET | Oxytocin, tetracycline | 342.72 | 0.2 g adsorbent L−1 | 130 mg g−1 for oxytocin, 150 mg g−1 for tetracycline | 0.38 mg m−2 for oxytocin, 0.44 mg m−2 for tetracycline | Langmuir | Pseudo-second-order | — | Chemisorption | 165 |
| Adsorption quantification: UV-vis | ||||||||||||
| Structural morphology: FE-SEM, TEM, AFM | ||||||||||||
| Others: XRD, FTIR, XPS, UPS, zeta potential | ||||||||||||
| MOF (MIL-101) | Solvothermal method | Surface: BET | Tetracycline hydrochloride | 180.41 | pH 10.2, 0.15 g adsorbent L−1, 10 mL L−1 H2O2 | 82.52% tetracycline hydrochloride removal efficiency | N/A | Pseudo-second-order | — | Photo–Fenton reaction | 166 | |
| Adsorption quantification: UV-vis | ||||||||||||
| Structural morphology: SEM, EDS | ||||||||||||
| Others: XRD, FTIR, XPS, zeta potential | ||||||||||||
| MOF (UiO-66-NH2) | Modified hydrothermal method | Surface: BET | Norfloxacin | 713.20 | At 0.10 g adsorbent L−1, pH 8 | 222.5 | 0.31 | Langmuir | Pseudo-second-order | — | Electrostatic, π–π and hydrophobic interaction | 167 |
| Structural morphology: SEM, TEM | ||||||||||||
| Others: XRD, FTIR, XPS, TGA | ||||||||||||
| MOF (Zn3(BTC)2) | Synthesis in a purely aqueous system | Surface: BET | Ofloxacin | At pH 7.7, 10 mg adsorbent L−1 | 25.3 ± 0.8 | Sigmoidal | Sigmoidal | — | N/A | 152 | ||
| Structural morphology: SEM | ||||||||||||
| Others: XRD, FTIR | ||||||||||||
| MOF (Zr/Fe-MOF/GO) | Hydrothermal method | Surface: BET | Tetracycline hydrochloride | At pH 6.7, 20 mg adsorbent L−1 | 156 | Freundlich | Pseudo-first order | — | Chemisorption | 168 | ||
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, TGA | ||||||||||||
| MOF (alginate-graphene-ZIF67) | Modified Hummers' method | Surface: BET | Tetracycline | 138.62 | At 30 °C, pH 6 | 456.62 | 3.29 | Freundlich | Pseudo-second-order | — | π–π interaction and cation-pi bond | 169 |
| Structural morphology: SEM | ||||||||||||
| Others: XRD, FTIR, XPS | ||||||||||||
| MOF (UiO-66-(COOH)2/GO) | Modified hydrothermal method | Surface: BET | Tetracycline hydrochloride | 369.60 | At universal pH, 0.50 g adsorbent L−1 | 164.91 | 0.45 | Langmuir | Pseudo-second-order | — | π–π interaction, chemical coordination, and weak electrostatic interaction | 170 |
| Structural morphology: SEM | ||||||||||||
| Others: XRD, FTIR, XPS, TGA | ||||||||||||
| MOF (UiO-66-NH-BT@g-C3N4) | Hydrothermal method | Surface: BET | Sulfamethoxazole | At pH 7, 25 mg PS dosage | 97.6% sulfamethoxazole removal efficiency | N/A | N/A | — | N/A | 171 | ||
| Structural morphology: SEM | ||||||||||||
| Others: XRD, XPS | ||||||||||||
| Graphene (graphite rods (15 cm in length and 1 cm in diameter)) | Electrochemical exfoliation method | Surface: BET | Di-n-butyl phthalate (DnBP), di-(2-ethylhexyl) phthalate (DEHP), acetaminophen (ACE), caffeine, cephalexin (CLX), and sulfamethoxazole (SMX) | At 0.10 g adsorbent L−1 and adsorption time of 12 h | 27.03, 39.22, 18.28, 22.73, 23.47, 17.42 respectively | Langmuir | Pseudo-second-order | — | Chemisorption | 172 | ||
| Structural morphology: TEM | ||||||||||||
| Graphene (NiZrAl-layered double hydroxide-graphene oxide-chitosan) | Hydrothermal method | Surface: BET | Nalidixic acid | 128.30 | At pH 8, 10 mg of adsorbent dosage | 277.79 | 2.17 | Temkin and Freundlich | Pseudo-second-order | — | Chemisorption | 173 |
| Structural morphology: SEM, TEM, EDX | ||||||||||||
| Others: XRD, RAMAN, pHZPC | ||||||||||||
| Graphene (graphene oxide (GO)) | Modified Hummers' method | Surface: BET | Tetracycline | At pH 3.6, 0.181 mg of adsorbent mL−1 | 313 | Langmuir and Temkin | Pseudo-second-order | — | Electrostatic interaction, π–π interaction, and cation–π bonding | 174 | ||
| Adsorption quantification: UV-vis, structural morphology: AFM | ||||||||||||
| Others: XRD, FTIR, RAMAN | ||||||||||||
| Graphene (graphene nanoparticles (GNP)) | Hydrothermal method | Surface: BET | Caffeine, acetaminophen, aspirin | 635 | At pH 8, 1 mg of adsorbent mL−1 | 18.76, 19.72, 13.02 | 0.03,0.031,0.02 | N/A | Pseudo-second-order | — | N/A | 175 |
| Adsorption quantification: HPLC-UV | ||||||||||||
| Structural morphology: TEM | ||||||||||||
| Others: XRD | ||||||||||||
| Graphene (reduced graphene oxide (rGO)) | Hydrothermal method | Surface: BET | Ciprofloxacin, norfloxacin | At pH 6.2, 2 mg of adsorbent dosage | 18.22–22.20 | Langmuir and Temkin | Pseudo-second-order | — | Electrostatic interaction, π–π interaction, and cation–π bonding | 176 | ||
| Structural morphology: SEM, TEM, EDX | ||||||||||||
| Others: XRD, FTIR, TGA | ||||||||||||
| Graphene (GO nanoparticles) | Hummers' method | Surface: BET | Metformin | 187.2 | At pH 6.26, (50–150) mg of adsorbent dosage | 122.61 | 0.66 | Freundlich | Pseudo-first order | −135.76, −2.15, 4.47 | Chemisorption, spontaneous, and endothermic adsorption process | 177 |
| Structural morphology: SEM, FE-SEM, TEM, EDX | ||||||||||||
| Others: XRD, FTIR, RAMAN, zeta potential, DLS, solid-state NMR | ||||||||||||
| Graphene (rGO) | Hydrothermal method | Surface: BET | Sodium diclofenac drug (s-DCF) | At pH 8–10, 30 mg of adsorbent dosage | 59.67 | Liu | General order | — | π–π interaction | 115 | ||
| Structural morphology: SEM, AFM | ||||||||||||
| Others: FTIR, RAMAN, TGA | ||||||||||||
| Graphene (graphene oxide composite reinforced with carboxymethyl cellulose) | Improved Hummers' method | Surface: BET | Amitriptyline (tricyclic antidepressant drug) | At pH 2–11, (2.5–40) mg of adsorbent dosage | 737.4 | Langmuir | Pseudo-second-order | — | Chemisorption and physisorption | 178 | ||
| Structural morphology: FE-SEM, EDX | ||||||||||||
| Others: FTIR, PZC, TGA | ||||||||||||
| Graphene (carbon xerogel/graphene hybrid) | Hydrothermal method | Surface: BET | Metronidazole (MNZ) | (648–816) | At pH 5 and 298 K | 110–166 | 0.17–20 | Freundlich–Langmuir | N/A | — | π–π and electrostatic interactions and hydrogen bonding | 179 |
| Structural morphology: SEM | ||||||||||||
| Others: XPS, RAMAN | ||||||||||||
| Biochar (hazelnut shell-based magnetic biochar) | Pyrolysis method | Surface: BET | Penicillin-G | At 40 °C, pH = 2, 1000 ppm Pen-G concentration, and 0.175 g of catalyst per 100 g of solution | 479 | Langmuir | Pseudo-second-order | — | Chemisorption | 120 | ||
| Structural morphology: SEM, EDX | ||||||||||||
| Others: FTIR, XRD, TGA | ||||||||||||
| Biochar (iron-loaded sludge biochar) | Pyrolysis method | Surface: BET | Tetracycline (TC) | 82.78 m2 g−1 | pH 2, contact time = 48 h, biochar dosage = 0.6 g L−1, initial TC concentration = 60 mg L−1 | 104.86 | 1.27 | Langmuir | Pseudo-second-order | — | π–π interaction, hydrogen bonds, complexation, and electrostatic interaction | 121 |
| Structural morphology: SEM, EDS, TEM | ||||||||||||
| Others: FTIR, XRD, XRF, XPS | ||||||||||||
| Biochar (sewage sludge-derived biochar) | Pyrolysis method | Surface: BET | Diclofenac (DCF) | 69.7–104.1 | At 500 °C, pH 3–6, contact time = 3 h, initial (DCF) concentration = 10–100 mg L−1 | 92.7 | Dubinin–Radushkevich | Pseudo-second-order | — | π-stacking interactions | 123 | |
| Biochar (maple leaf-derived biochar) | Pyrolysis method | Surface: BET | Tetracycline (TC) | 191.1 | pH 6–9, contact time = 5 days, biosorbent dosage = 0.01 g L−1, initial TC concentration = 100 mg L−1 | 407.3 | 2.13 | Freundlich | Pseudo-second-order | — | Metal complexation, H-bonding, and hydrophobic interactions | 124 |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, XPS, TGA, ICP-OES, pHPZC | ||||||||||||
| Biochar (algal-based (spirulina species) biochar) | Pyrolysis method | Surface: BET | Tetracycline (TC) | 1.55 (specific surface area) | At 550 °C, pH 6, contact time = 48 h, biochar dosage = 0.05 g L−1, initial TC concentration = 100 mg L−1 | 132.8 | Langmuir | Pseudo-first order | — | Hydrophobic, π–π interactions, functional groups, and metal complexation | 125 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, XPS, TGA, ICP-OES, pHPZC, elemental analyzer | ||||||||||||
| Biochar (walnut shell biochar) | Pyrolysis method | Surface: BET | Sulfadiazine | At 25 °C, pH 1–10, contact time = 48 h, biochar dosage = 0.02 g L−1, initial sulfadiazine concentration = 20 mg L−1 | 32 | Freundlich | Elovich | — | π–π EDA interaction, hydrogen bond, electrostatic interaction, Lewis's acid–base interaction, and hydrophobic interaction | 126 | ||
| Adsorption quantification: HPLC-UV | ||||||||||||
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, XPS, RAMAN, elemental analyzer | ||||||||||||
| Biochar (sugarcane bagasse-derived biochar) | Pyrolysis method | Surface: BET | Sulfamethoxazole | 1099 | At 303 K, pH 2–10, contact time = 80–90 min, biochar dosage = 0.05 g L−1, initial sulfadiazine concentration = 100 mg L−1 | 400 | 0.37 | Freundlich | Elovich | — | π–π interaction and hydrogen bonding | 127 |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: FTIR, XRD, pHPZC | ||||||||||||
| Biochar (peanut shell-derived biochar) | Pyrolysis method | Surface: BET | Naproxen | 596 | At 25 °C, pH 7, contact time = 48 h, biochar dosage = 0.5 g L−1, initial naproxen concentration = 1000 mg L−1 | 324 | Langmuir | Pseudo-second-order | −20.3, −18.8, 4.47 | Pore filling, π–π interaction, hydrogen bonding formations, n–π interaction, van der Waals force, and electrostatic attraction | 129 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, pHPZC, RAMAN | ||||||||||||
| Biochar (bamboo sawdust) | Pyrolysis method | Surface: BET | Acetaminophen | 1158.05 | At 25 °C, pH 6.8, contact time = 120 min, biochar dosage = 0.5 g L−1, initial acetaminophen concentration = 20 mg L−1 | 192.43 | Langmuir | Pseudo-second-order | — | Chemisorption | 180 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, RAMAN | ||||||||||||
| Biochar (bamboo sawdust) | Pyrolysis method | Surface: BET | Ciprofloxacin | 1158.05 | At 25 °C, pH 6.8, contact time = 120 min, biochar dosage = 0.5 g L−1, initial ciprofloxacin concentration = 20 mg L−1 | 70.95 | Langmuir | Pseudo-second-order | — | Chemisorption | 180 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, RAMAN | ||||||||||||
| Biochar (corn-cob-derived biochar) | Pyrolysis method | Surface: BET | Acetaminophen | 1201.1 | At 20–40 °C, pH 2–12, contact time = 0–350 min, biochar dosage = 1 g L−1, initial acetaminophen concentration = 0–500 mg L−1 | 332.08 | Langmuir | Pseudo-second-order | −8.58, −44.39, −0.12 | Hydrogen bonding formation and n–π interactions | 181 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, RAMAN | ||||||||||||
| Biochar (corn-cob-derived biochar) | Pyrolysis method | Surface: BET | Amoxicillin | 1201.1 | At 20–40 °C, pH 2–12, contact time = 0–360 min, biochar dosage = 1 g L−1, initial amoxicillin concentration = 0–500 mg L−1 | 175.86 | Freundlich | Pseudo-first order | −1.02, −4.68, −0.01 | Hydrogen bonding formation and n–π interactions | 181 | |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, RAMAN | ||||||||||||
| Biochar (Prosopis juliflora) | Pyrolysis method | Surface: BET | Sulfamethoxazole | 875.8 | At 25–50 °C, pH 8, contact time = 120 min, biochar dosage = 1 g L−1, initial sulfamethoxazole concentration = 50 mg L−1 | 49.776 | Langmuir | Pseudo-second-order | — | Chemisorption | 182 | |
| Structural morphology: SEM, EDX | ||||||||||||
| Others: FTIR, XRD, pHPZC, proximate analysis, TGA | ||||||||||||
| Biochar (Prosopis juliflora) | Pyrolysis method | Surface: BET | Ciprofloxacin | 875.8 | At 25–50 °C, pH 5, contact time = 120 min, biochar dosage = 1 g L−1, initial ciprofloxacin concentration = 50 mg L−1 | 91.432 | Langmuir | Pseudo-second-order | — | Chemisorption | 182 | |
| Structural morphology: SEM, EDX | ||||||||||||
| Others: FTIR, XRD, pHPZC, proximate analysis, TGA | ||||||||||||
| Agricultural waste-based materials (coconut shell-derived activated carbon) | Pyrolysis method | Surface: BET | Levodopa | 1175 | At 25 °C, pH 6.2, contact time = 120 min, initial dosage = 0.1 g L−1, initial levodopa concentration = 0.013 mg L−1 | 285.3 | 0.24 | Freundlich | N/A | — | Donor–acceptor mechanism | 183 |
| Adsorption quantification: UV-vis | ||||||||||||
| Others: TGA, pHPZC | ||||||||||||
| Agricultural waste-based materials (carbon foam pellets derived from Vallisneria natans) | Ball milling, pyrolysis method, hydrothermal method | Surface: BET | Metronidazole | 922.56 | At 30 °C, pH 9.0, contact time = 120 min, initial dosage = 6.0 g L−1, initial metridazole concentration = 10 mg L−1 | 64.23 | 0.069 | Langmuir | Pseudo-first order | — | Hydrogen bonding, π–π interactions, and micropore filling | 184 |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, pHPZC, elemental analyzer, TGA | ||||||||||||
| Agricultural waste-based materials (grapestalk-derived activated carbon) | Knife milling, hydrothermal method | Surface: BET | Caffeine | 1099.86 | At 30 °C, pH 4, contact time = 30 min, initial dosage = 15.0 g L−1, initial caffeine concentration = 5–35 mg L−1 | 916.7 | 0.83 | Sips | N/A | — | N/A | 133 |
| Others: pHPZC | ||||||||||||
| Agricultural waste-based materials (iron(III)-loaded bamboo cellulose nanofibers) | Mechanical shearing method, freeze-dryer | Surface: BET | Tetracycline (TC) | 171 | At 25 °C, pH 7, contact time = 30 min, initial adsorbent dosage = 0.5 mg L−1, initial caffeine concentration = 5–10 mg L−1 | 294.12 | 1.72 | Langmuir | Pseudo-second-order | — | Surface complexation, hydrogen bonding, electrostatic interaction, and van der Waals force | 185 |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: FTIR, XRD, pHPZC, XPS, TGA | ||||||||||||
| Agricultural waste-based materials (iron(III)-loaded bamboo cellulose nanofibers) | Mechanical shearing method, freeze-dryer | Surface: BET | Chlortetracycline | 171 | At 25 °C, pH 7, contact time = 30 min, initial adsorbent dosage = 0.5 mg L−1, initial caffeine concentration = 5–10 mg L−1 | 232.56 | 1.36 | Langmuir | Pseudo-second-order | — | Surface complexation, hydrogen bonding, electrostatic interaction, and van der Waals force | 185 |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: FTIR, XRD, pHPZC, XPS, TGA | ||||||||||||
| Agricultural waste-based materials (activated carbon derived from olive stones) | Pyrolysis method | Surface: BET | Amoxicillin | 1174.00 | At 25 °C, pH 7, contact time = 4000 min, initial adsorbent dosage = 1 g L−1, initial caffeine concentration = 12.5–100 mg L−1 | 67.7 | 0.058 | Sips | Pseudo-second-order | — | Chemisorption | 186 |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: pHPZC | ||||||||||||
| Naturally occurring materials (chitosan-grafted SiO2–Fe3O4) | Co-precipitation method | Surface: BET | Ciprofloxacin | 126.16 | At 37 °C, pH 7.4, contact time = 2 h, initial adsorbent dosage = 1–10 mg L−1, initial ciprofloxacin concentration = 12.5–10 mgL−1 | 100.74 | 0.79 | Langmuir | Pseudo-second-order | — | Monolayer mechanism | 136 |
| Adsorption quantification: UV-vis | ||||||||||||
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, XRD, zeta potential, TGA, vibrating sample magnetometer | ||||||||||||
| Naturally occurring materials (NiFe2O4–COF-chitosan-terephthalaldehyde nanocomposites film) | Hydrothermal method | Surface: BET | Tetracycline | 107.33 | At 25 °C, pH 3–11, contact time = 30 min, initial adsorbent dosage = 5.4 mg, initial tetracycline concentration = 10 to 400 mg L−1 | 388.52 | 3.62 | Freundlich | Pseudo-second-order | — | Complexation, cation exchange, electrostatic attraction, hydrogen bonding, and the π–π interaction | 137 |
| Structural morphology: SEM, TEM, HR-TEM, EDX | ||||||||||||
| Others: FTIR, XRD, TGA, elemental analysis | ||||||||||||
| Naturally occurring materials (magnetically engineered sulfurized peat-based activated carbon) | Pyrolysis method | Surface: BET | Sulfamethoxazole (SMX) | 724 | At 311 K, pH 11, contact time = 30 min, initial adsorbent dosage = 5.4 mg L−1, initial sulfamethoxazole (SMX) concentration = 1.4 mg L−1 | 94% | Langmuir (linear) and Freundlich (non-linear) isotherms | Pseudo-second-order | — | π–π electron donor–acceptor interactions, hydrogen bonding | 138 | |
| Structural morphology: SEM, EDS | ||||||||||||
| Others: FTIR, XPS, XRD, zeta potential | ||||||||||||
| Naturally occurring materials (bark-based biochar) | Pyrolysis method | Surface: BET | Tetracycline | 683.33 | At 10–50 °C, pH 2–12, contact time = 10–120 min, initial adsorbent dosage = 1–2 g L−1, initial tetracycline concentration = 50–500 mg L−1 | 58.47 | 0.085 | Langmuir | Pseudo-second-order | — | Chemisorption | 187 |
| Structural morphology: SEM | ||||||||||||
| Others: FTIR, pHPZC |
![]() | ||
| Fig. 8 Biochar production and modification using various sources for the adsorption and elimination of pharmaceutical contaminants from wastewater (adapted from ref. 119 with permission from Elsevier, copyright 2024). | ||
The highest concentrations of pharmaceutical pollutants recovered with biochar include tetracycline (found at 1163 mg g−1), sulfamethoxazole (found at 400 mg g−1), naproxen (596 mg g−1), and norfloxacin (698.6 mg g−1) when using biochar derived from corncobs.118 Unlike other adsorbents, biochar can be recycled up to eight times with minimal efficiency loss.118 Additionally, hazelnut shell was utilized as a precursor in producing magnetic biochar (MBC), which was subsequently applied as a biosorbent to eliminate pharmaceutical impurities from contaminated water. This hazelnut shell biochar achieved the highest Pen-G adsorption capacity of 479 mg g−1 at 40 °C, pH 2, 1000 ppm penicillin-G (Pen-G) concentration, and 0.175 g of adsorbent per 100 g of solution.120 Iron-loaded sludge biochar showed excellent performance, with a surface area of 82.78 m2 g−1, removing tetracycline (TC) to a maximum adsorption capacity of 104.86 mg g−1 under pH levels of 2–10.121 Furthermore, biosolid (mostly biowaste)-derived biochar has an excellent surface area of approximately 182 m2 g−1 and removed triclosan from wastewater with a capacity of about 1330 µg g−1 at pH 7.122 Sewage sludge-derived biochar also showed excellent potential for wastewater treatment as it removed diclofenac (DCF), naproxen (NAP), and triclosan (TCS) with maximum adsorption capacities of 92.7 mg g−1, 127 mg g−1, and 113 mg g−1, respectively, from polluted water at pH 2–11.123 Biochar generated from maple leaves at three temperatures (350 °C, 550 °C, and 750 °C) produced the highest sorption rate for tetracycline, with an adsorption capacity of 407.3 mg g−1 at a pyrolytic temperature of 750 °C.124 Algal-based (Spirulina species) biochar generated at a temperature of 750 °C was shown to be more effective in treating tetracycline waste than biochar generated at 350 °C and 550 °C, with an adsorption potential of 132.8 mg g−1.125 Sulfadiazine, sulfamethazine, and sulfachloropyridazine can all be removed by walnut shell biochar with removal efficiencies of 32 mg g−1, 46 mg g−1, and 40 mg g−1, respectively.126 Sulfamethoxazole was significantly removed from biochar made from sugarcane bagasse by hydrothermal carbonization at 200 °C and alkali activation with NaOH in an inert environment. A maximal sorption capacity of 400 mg g−1 was reported by Prasannamedha et al.127 Novel biochar derived from corn stalk, reed stalk, and willow branches was studied by Wang et al.,128 and used for norfloxacin removal in water, achieving maximum adsorption capacities of 7.2469, 3.5139, and 6.2587 mg g−1, respectively. Moreover, peanut shell-derived biochar, which was prepared by pyrolysis at 800 °C for approximately 4 hours, removed naproxen with an excellent adsorption capacity of 324 mg g−1,.129
| Adsorbent category | Cost of material | Regeneration potential | Carbon footprint (kg CO2 eq. kg−1 adsorbent) | E-factor (kg of waste/kg of desired product) | Lifecycle assessment outcome | Techno-economic analysis based on capital expenditure (CAPEX) and operating expenditure (OPEX) |
|---|---|---|---|---|---|---|
| Metal–organic frameworks | $20–50 kg−1 (ref. 197 and 198) | Typically limited (<5 cycles); stability issues199 | 23.7–2482 (ref. 188) | 1.3–19.9 (ref. 200) | Very high, as the carbon footprint is substantial | Scaling is limited due to very high CAPEX and OPEX197,198 |
| Graphene/graphene oxides | $10–25 kg−1 (ref. 201) | Moderate; potential loss due to oxidation202 | 60–594 (ref. 188) | 0.1–2.5 (ref. 203) | High due to its high global warming potential/carbon footprint | Moderately scalable as OPEX is moderate201 |
| Zeolites | $2–4 kg−1 (ref. 204) | Good, thermally stable205 | 1–15 (ref. 189) | 21 (ref. 206) | Moderate due to its moderate carbon footprint | Highly scalable as it requires low CAPEX and OPEX204 |
| Metal oxide nanoparticles | $5–50 kg−1 (ref. 116) | Typically, 5–10 cycles with >85–95% capacity retention207 | 9.9 × 10−2–3.7 (ref. 208) | 2.1–15 (ref. 209) | Lower due to its low carbon footprint value | OPEX is moderate, and scaling is moderate116 |
| Biochar | Less than $1 kg−1 (ref. 210) | 5–8 cycles depending on surface functionalization211 | −6.3–1 (ref. 212) | 1.4 (ref. 213) | Negative as these are mainly produced from carbon sources and possess low end-of-life risk | Highly scalable due to having the lowest CAPEX and OPEX values compared to the other adsorbent categories210 |
After removing pollutants, adsorbents with high aquatic stability may readily separate from wastewater streams. The reusability of used adsorbents is determined by their capacity for recovery, decontamination, and regeneration. The adsorbent needs to be reused and regenerated to reduce costs for industrial applications. The regeneration method must be chosen carefully to effectively desorb the pollutant. The viability of industrial-scale use depends on several factors, including the kind of adsorbent, the contaminants, the adsorbent's stability, the toxicity of the used adsorbents, and the cost and energy requirements of the regeneration process. Filtration, magnetic separation, thermal desorption, microwave irradiation, advanced oxidation process, solvent regeneration, and microbial-assisted adsorbent regeneration are several techniques for regenerating spent adsorbent. A bar magnet can readily separate magnetic biochar from biomass that has been pre-treated with iron salts like K2Fe2O4 and FeCl2/FeCl3 to create magnetic biochar.190 An adsorbent created by coating palygorskite with magnetite nanoparticles displayed a magnetic susceptibility of 20.2 emu g−1 and absorbed 26.6 mg g−1 of Pb2+ from water. The spent adsorbent was easily separated using a basic bar magnet.191 An et al. showed the potential for excellent sorbent reusability.192 Up to the fifth run, the number of MOF-derived carbons (MDCs) required to remove pharmaceutical products from water did not significantly grow as the number of cycles increased. Moreover, the performance was still around an order of magnitude higher than the brand-new activated carbon (AC) after the fifth run. Furthermore, metal recovery by thermal desorption from used adsorbent is an emerging method. MWCNT (multi-walled carbon nanotubes) were successfully recycled by Toński et al. by thermal desorption, and effectively used to remove cyclophosphamide, ifosfamide, and 5-fluorouracil.193 Using 0.5 M NaOH, arsenic could be desorbed from magnetic sorbents, and additional magnetic adsorbents could be regenerated, as demonstrated by Baig et al.194 When used as regeneration solvents, HCl, HNO3, and H2SO4 showed considerable desorption efficiency.195 Advanced oxidation processes (AOPs) for regenerating used adsorbents have gained popularity in recent years. Yang et al. studied the bio-regeneration of clays or functionalized clays and reported that microbial regeneration of montmorillonite functionalized with hexadecyltrimethylammonium (HDTMA) was superior to chemical regeneration.196
Environmental and societal problems might result from the open disposal of used adsorbents that contain harmful organic pollutants, especially in developing countries with limited access to designed landfills and incinerators. That is why the management of spent sorbents is significant. Although disposal is a cost-effective method, it is crucial to consider its environmental viability and long-term sustainability. Four methods (reuse, regeneration, repurposing/recycling, and final disposal) have been utilized for the sorbent's management, disposal, and repurposing. While landfilling and incineration are standard safe disposal methods, reused waste adsorbents are employed in applications including soil amendment, capacitors, and catalyst/catalyst support.
A synergistic effect on reproduction in natural surface water containing a quaternary mixture of FLU, metformin, ASA, and CIP was reported by Tominaga et al.219 In a mixture, the pollutants can exhibit independent mechanisms (independent action model) or a common mechanism (all pollutants show the same effect with a precise mechanism, concentration action model). Hence, more studies should focus on the ecotoxicological effects of pharmaceutical pollutants in mixtures. It emphasizes the importance of conducting ecotoxicological studies under more realistic conditions and considering the interactions between PWs. Furthermore, in most studies, the treatment process and performance have been highlighted, while the toxicity assessment of the treated wastewater has been overlooked. The treated water may still be ecologically toxic. Furthermore, the possibility of secondary pollution by the adsorbent (e.g., metal ion leaching from perovskites) and transformed products generated during the treatment process should not be overlooked. Research related to PW treatment should include toxicity tests to validate the performance of the proposed treatment method.
Over the last two and a half decades, numerous articles have been published on the removal of pharmaceutical pollutants through adsorption. However, less than 5% of these works have investigated competitive adsorption.220 In the practical environment, various pharmaceutical pollutants are expected to be present in wastewater at varying concentrations. Investigators should consider this reality in future projects. This mixture of different pollutants also presents a challenge in designing adsorbents. Since different active ingredients contain distinct functional groups in their structures, the adsorption efficacy regarding a single target compound may vary depending on the adsorbent. The synthesis of green materials containing various functional groups that can adsorb multiple pollutants can be an interesting topic to explore in the future. At the same time, researchers should focus on performing regeneration studies of these materials for waste minimization and cost optimization.
Since pharmaceutical pollutants (PPs) are emerging contaminants, existing wastewater treatment plants may have lower pollutant removal efficiencies. However, utilizing the existing treatment infrastructure and modifying these already established facilities will be more economical and time-consuming. Integrating multiple processes with the existing ones can improve performance.221,222 Adsorption of IBU, CAR, and 17 α-ethinyl estradiol (EE2) using activated biochar (ABC) followed by an ultrafiltration (UF) membrane increased the retention rate to 41.8%, 40.9%, and 53.0% from 24.4%, 7.0%, and 14.8%, respectively, in a UF alone system without decreasing the flux rate considerably (normalized flux rate in ABC = 0.85).215 Approximately 10% more COD removal from a pharmaceutical industrial effluent containing anti-psychotic and anti-cancer ingredients was achieved using advanced oxidation (ozone + peroxide)-activated char treatment (85.4%) than oxidation alone (75%) at pH 5.223 A combined activated sludge–activated carbon system removed 100% (2 mg L−1 each) of acetaminophen, IBU, and caffeine, showing better results than the biological treatment alone. More studies should focus on hybrid treatment methods for efficiently removing pharmaceutical components from wastewater.224
Although adsorption has been reported as an efficient method for PP removal from wastewater, laboratory results do not represent the practical pilot-scale/actual performance of the treatment process. Since wastewater has a highly complex character, accurately simulating wastewater in the lab is rarely attainable. It hinders the evaluation of the actual performance of the adsorbent. Additionally, further adsorption studies of various materials, such as perovskites, should be conducted in practical settings. The performance and cost of treatments performed in a laboratory setting can differ significantly in real-life applications. Further pilot-scale studies are needed to develop more robust, efficient, and cost-effective treatment methods.
Techno-economic analysis (TEA) is a crucial tool for evaluating the feasibility of industrial processes. TEA of the adsorptive treatment of pharmaceutical wastewater is scarce. Echevarría et al.225 performed TEA on an advanced water reclamation pilot plant operating at a capacity of 1.5–2 m3 h−1. Two ultrafiltration-reverse osmosis (UF-RO) blends, including only RO and a powdered activated carbon (PAC)-tight UF, were evaluated for treating wastewater containing CAR, DIC, ERY, SUL, and diuron. PAC-tight UF showed 81 ± 13% removal efficiency, while 55 ± 11% pollutants were removed by UF-RO (50% blend). UF-RO (50%) required the lowest operating cost at €0.18 m−3, followed by PAC-tight UF (€0.22 m−3; 20 mg L−1 PAC), 25% UF-RO (€0.24 m−3), and OR (€0.31 m−3). The lowest capital expenditure of €548 m−3 was estimated for PAC-tight UF, while 50% UF-RO, 25% UF-RO, and RO would cost €594 m−3, €628 m−3, and €662 m−3, respectively. As the ecological threat of PPs is mounting, the necessity of more techno–economic studies in this regard has become paramount. The techno–economic feasibility of resource recovery (e.g., pharmaceutical precursors) from PW can be a predominant research direction to ensure the robustness of future treatment plants.
Footnote |
| † Co-first authors. |
| This journal is © The Royal Society of Chemistry 2025 |