Conventional treatments for antibiotic residues in effluents are inefficient and do not lead to complete removal. Though effective and feasible degradation of antibiotics using nanoparticles has been reported by several scientists, chemically synthesized nanoparticles have their own disadvantages. Thus, in this study, nZVI was biosynthesized using leaf extract of Shorea robusta and precursor FeSO4·7H2O for photocatalytically degrading tetracycline (TC) and ciprofloxacin (CIP). The characterization of nZVI was performed using SEM, TEM, AFM, EDX, FTIR, and XRD to test their properties, which revealed iron-rich, well-dispersed, spherical, crystalline nanoparticles. Photocatalytic degradation of TC and CIP under UV illumination revealed 88 and 84% optimum efficiency at antibiotic concentrations 15 and 25 mg L−1, 0.014 and 0.0175 g L−1 doses of nZVI, respectively in the pH range 4–6 in 70 min. The degradation was further verified using mass spectrometry, which confirmed the degradation of antibiotics into the breakdown products. Toxicity assay of the degraded antibiotic solution proved it non-toxic for bacteria and safe for discharge into water bodies. The cost analysis of antibiotic degradation using nZVI proved very economical, costing around 1.5 USD per 1,000 L of wastewater.

  • nZVI was synthesized by leaf extract of Shorea robusta and FeSO4 · 7H2O as the precursor.

  • Analytical technique confirmed the characteristics of well-formed nZVI.

  • Synthesized nZVI in the presence of UV light could degrade 88% of TC and 84% of CIP.

  • Mass spectroscopy and toxicity assay indicated the breakdown of antibiotics into smaller, non-toxic compounds.

  • Application of nZVI for the treatment of TC and CIP proved to be very inexpensive.

Graphical Abstract

Graphical Abstract
Graphical Abstract

Antibiotics are chemicals active against bacterial infections and are used in human medication, veterinary medication, aquaculture additives, poultry industry additives, and agriculture growth (Le et al. 2018; Ben et al. 2019). A significant portion of these antibiotics remain unmetabolized and are excreted into the environment. Antibiotic residues are persistent and adversely affect the environment and human health (Al Omari et al. 2019; Ben et al. 2019). Increased application of antibiotics has led to increased antibiotic residues in surface water, groundwater, wastewater and soil, and their concentration of up to a few μg L−1 has been reported in the natural water system. Antibiotic residue in water can form antibiotic-resistant bacteria and antibiotic-resistant genes through mutational or genetic changes, death of sensitive microorganisms, and altering the microbiome even at low concentrations. This can further lead to antibiotic resistance, treatment failure, and mortality in humans. Antibiotic residues in wastewater are toxic to algae and aquatic animals and are potential mutagens and carcinogens (Tan et al. 2015; Ezzariai et al. 2018).

Tetracycline (TC) is a broad-spectrum antibiotic used in the treatment of gram-negative, gram-positive bacteria, fungus, and other microorganisms (Hou et al. 2020). TC is one of the most used antibiotics worldwide, has a long half-life, and is non-biodegradable and highly soluble in water. TC is extensively used for medical treatment, livestock, and aquaculture. TC is frequently detected in groundwater, surface water, wastewater, and soil and can form different byproducts depending on pH and other environmental conditions. The toxic effect of TC has been reported on plants, sewage sludge bacteria, and non-target organisms (Guo et al. 2017). TC also acts as a chelating agent for metal ions, forming complexes without reducing their negative impacts in an aqueous solution.

Ciprofloxacin (CIP) is the second generation of fluoroquinolone antibiotics, effective against gram-negative and some gram-positive bacteria. CIP is an extensively used antibiotic found in high concentrations in hospital wastewater and wastewater treatment plants, up to a few mg L−1 (Nawaz et al. 2020). CIP and its residues can have an eco-toxic effect and lead to the formation of antibiotic-resistant genes (Orimolade et al. 2020).

Conventional treatment plants are insufficient in removing antibiotic residues from wastewater, resulting in their penetration into receiving water bodies (Guo et al. 2017; Orimolade et al. 2020). Antibiotics have been successfully treated by the application of nanotechnology in recent times. Nanoparticles are materials with dimensions less than 100 nm (Li et al. 2021). Nanoparticles can be zero-, one-, two, or three-dimensional based on their shape. Nanoparticles generally have three distinct layers; the core, shell, and surface.

nZVI can be synthesized by the bottom-up or top-down method (Stefaniuk et al. 2016). The top-down approach utilizes the diminution and processing of the precursor to produce nanoparticles, while the bottom-up process uses simpler molecules to form nZVI. Recent technologies of nZVI synthesis include precision milling, carbothermal method, ultrasound-assisted method, electrochemical method, and green synthesis (Stefaniuk et al. 2016). Precision milling works by milling the iron to produce particles in the nano-size range, but their shapes are irregular. The carbothermal process uses Fe2+ salts, elevated temperature (>500 °C), and H2, CO2, or CO as reducing agents. Ultrasound can be combined with chemical methods to produce significantly smaller particles, but this process can simultaneously oxidize the nZVI. The electrochemical method utilizes Fe2+/Fe3+ salts in solution, and nZVI accumulates at the cathode, but this method is affected by aggregation. Green synthesis is the most environmentally friendly method of nZVI synthesis, eliminating the use of harmful chemicals and high temperatures, and is economical, with a possibility of large-scale synthesis (Jha & Chakraborty 2020). In an aqueous solution, nZVI produces highly reactive species such as superoxide ions, electrons, hydroxide radicles, and hydroxide ions, which helps in the degradation of organic pollutants (Jha & Chakraborty 2020). nZVI has proven to be highly efficient in the treatment of environmental pollutants. nZVI can remove pollutants through adsorption, oxidative degradation, precipitation, and reductive degradation (Ezzatahmadi et al. 2017). The hypothesis of this study lies in the fact that nZVI corrodes to form Fe2+, which activates H2O2 to form hydroxyl radicals and superoxide ion, having great oxidizing capabilities toward degradation of antibiotics by hydroxylation, ring opening, and fragmentation (Li & Liu 2021). Zhao et al. (2020) synthesized zeolite-supported nZVI for the removal of Norfloxacin and Ofloxacin at a concentration of 10 mg L−1 with an efficiency of 90% in 1 h and a maximum uptake-capacity of 54.67 and 48.88 mg g−1, respectively. Leili et al. (2018) performed adsorption of cephalexin by nZVI with an uptake-capacity of 1,667 mg g−1 and an efficiency of 83.8% at pH 2 in 2 h. Tran et al. (2020) applied 0.5 g L−1 platinum/nZVI composite for 100% degradation of oxytetracycline in 20 min at pH 5. Hou et al. (2020) successfully utilized nZVI loaded metal-organic framework to treat TC, with an efficiency of 90% and an uptake-capacity of 625 mg g−1 in a broad pH range. A search was conducted on the web of science (https://www.webofscience.com/wos/woscc/basic-search) on the 26th of February 2022, with the topic ‘antibioti*nZVI’. The search resulted in 101 articles, including 97 research articles and 4 review articles. Results were filtered for research articles and shorted by the highest citation first. The first article was published in 2009. The top 10 most cited articles are provided in Supplementary Table S1, and the top 5 most cited authors, organizations, countries, publishers, journals, and research areas are provided in Supplementary Table S2. The literature review from the web of science indicated that most of the research on removal of antibiotics using nZVI composites was based on the adsorption mechanism. Composites of nZVI and activated carbon, biochar, carbon nanotubes, chitosan, clay, graphene oxide, reduced graphene oxide, silica, and zeolite have been successfully used as an adsorbent. While for degradation studies, bimetallic composites of TiO2, Ni, and Cu have been applied for degradation of antibiotics. Moreover, most of the reported degradation processes are assisted by fenton, sonocatalysis, flocculation, or membrane. Reports on degradation on antibiotics using only nZVI are limited. These processes incur more cost and incomplete solution for treatment of antibiotics. Thus, in our study, we have attempted application of biosynthesized nZVI, for treatment and complete degradation of TC and CIP in the presence of UV light.

Experimental design

Biosynthesis of nZVI was achieved by the reduction of Ferrous Sulfate Heptahydrate (FeSO4·7H2O) by extract of matured leaves from Shorea robusta (Sal). The process flow diagram of the work design is shown in Figure 1.
Figure 1

Experimental design for application of biosynthesized nZVI for degradation of TC and CIP.

Figure 1

Experimental design for application of biosynthesized nZVI for degradation of TC and CIP.

Close modal

Reagents used

The two antibiotics, ciprofloxacin (C17H18FN3O3) and tetracycline (C22H24N2O8), were obtained from Alfa Aesar, ferrous sulfate heptahydrate (FeSO4·7H2O), which was used as the precursor for nZVI, was obtained from Merck, and sulfuric acid (H2SO4) and sodium hydroxide (NaOH) were obtained from Rankem. Further work was performed by Milli-Q water, and all the chemicals were used without further purification.

Preparation of antibiotic solutions

Stock solutions of CIP (500 mg L−1) and TC (500 mg L−1) were prepared by dissolving 500 mg of CIP and TC in 1,000 ml of Milli-Q water separately and mixed homogeneously using an ultrasonic homogenizer (Aczet APU 500). These were diluted to 100 mg L−1, and the working solutions (10–35 mg L−1) were prepared by successive dilution with Milli-Q water. The concentration range was selected based on the fact that the maximum CIP concentration of up to 31 mg L−1 was reported in raw wastewater of the Indian subcontinent by Bhagat et al. (2020) and the TC concentration of up to 20 mg L−1 has been reported from animal manure (Winckler & Grafe 2001). Globally, the concentration of antibiotics in hospitals and urban wastewater has been reported to be in the range of 0.3–100 mg L−1 (Abbasnia et al. 2022; Parmar & Srivastava 2022).

Biosynthesis of nZVI

nZVI was biosynthesized according to the previous method (Jha & Chakraborty 2020) with slight modification. Leaf extract was used as the reducing agent for the precursor, FeSO4·7H2O. For this, matured leaves from Shorea robusta (Sal) were plucked from BIT Mesra campus, Ranchi, Jharkhand, India (23.418359°N, 85.438360°E) and rinsed with Milli-Q water to remove adhered impurities, blot-dried, and air-dried, respectively. 20 g of leaves were weighed and crushed using mortar and pestle with water to a fine paste, and the final volume of water was made to 200 ml (1 g of leaf per 10 ml of water) and finally filtered using a membrane filter (pore size 45 μm). For future consumption, the filtrate was stored at 4 °C.

To synthesize nZVI, 50 ml of 0.01 M FeSO4·7H2O solution was taken in Erlenmeyer flasks, and 50 ml of the Sal leaf extract was added dropwise at room temperature and placed in an orbital shaker at 100 rpm. The formed particles were characterized by various methods. The particles were separated from the liquid phase by the vacuum filtration technique, washed with absolute ethanol and stored submerged in ethanol to avoid oxidation of nZVI.

Characterization of biosynthesized nZVI

The general characterization of the formed particles was performed using various techniques. The morphology of the formed particles was determined using transmission electron microscopy (TEM; CM 200; operating voltage: 20–200 kV) and atomic force microscopy (AFM; NT-MDT, Russia; Solver Pro-4). Particle size was determined using light scattering (Make: Malvern Instruments, UK, Nano ZS, 4.0 mW, 633 nm) and TEM images. Energy-dispersive X-ray spectroscopy (Make: JOEL JSM-6390LV) was employed to determine the elemental composition. X-ray diffraction (XRD; Rigaku, Japan, SmartLab 9 kW) was used to confirm the synthesis of nZVI. The determination of the functional groups in the Sal leaf extract and the synthesized particles was done using Fourier transform infrared spectroscopy (FTIR; Make: Shimadzu Corporation, Japan, IR-Prestige 21, Range: 4,000–400 cm−1). FTIR indicated the probable functional groups responsible for the reduction of FeSO4·7H2O and the functional groups which could help in the degradation of antibiotics.

Degradation study of TC and CIP

The degradation of TC and CIP solutions was studied in batch culture in triplicates under various experimental conditions and monitored using a UV–Visible spectrophotometer (Shimadzu corp UV-1800). The current study recorded the peak of TC and CIP at wavelengths 357 and 277 nm (Che et al. 2018). 50 ml of TC and CIP solutions were taken in 100 ml Erlenmeyer flasks, nZVI was added to it, and placed in UV light (2 lights of Philips TL-D 18 W each) in a closed chamber (Size 30 cm*30 cm*45 cm) under magnetic stirring at 100 rpm (Tarsons, Spinot Digital Model MC-02). Initially, the concentration of TC and CIP was kept at 10 mg L−1, and the dose of nZVI at 0.01 g L−1, for 1 h. Degradation efficiency was calculated based on Equation (1) (Selvamani et al. 2021)
(1)
where C0 is the initial concentration and Ct is the concentration of antibiotic after time t.

To study the effects of light, control experiments were set up in dark and sunlight. Initial concentrations of TC and CIP were varied as 10, 15, 20, 25, 30, and 35 mg L−1; pH as 2, 4, 6, 8, 10, and 12. The adjustment in the pH of antibiotic solutions was carried out using HCl (0.1 M) and NaOH (0.1 M). Doses of nZVI varied as 0.0035, 0.0070, 0.0105, 0.0140, 0.0175, and 0.0210 g L−1. The stirring rate was varied as 100, 200, 300, 400, and 500 rpm, and time was optimized up to 70 min with equal intervals of 10 min. Mass spectrometry (Thermo Scientific LTQ-XL) of the degraded antibiotics solutions was performed in the range of 150–500 m/z to determine the degraded products.

Microbial assay

To study the toxic effect of photocatalytically degraded TC and CIP solution on microorganisms, a bioassay was performed by the cup plate technique. The water sample was collected from the Subarnarekha River (23.4071275°N, 85.4390985°E). Bacterial strains were isolated through serial dilution and plating on a nutrient agar plate. Two bacterial strains were isolated and sub-cultured using nutrient broth media based on the distinct colors. Gram staining, as well as bioassay, was performed by the standard method. Three cups were bored, marked as A, B, and C, and filled with initial concentrations of antibiotics, degraded antibiotic solutions, and Millipore water, respectively, and incubated at 37 °C. After 24 h, the zone of inhibition was measured to study the impact on bacterial strain.

The zone of inhibition was recorded after incubation as an indication of the toxicity of the bacterial strains.

Synthesis and characterization of nZVI

Addition of Sal leaf extract to the precursor (FeSO4·7H2O) in the ratio of 1:1 led to the change in color of the solution from faint yellow to black, indicating the successful bio-reduction of FeSO4·7H2O to nZVI. The UV–Visible spectrum of the synthesized material confirmed the synthesis of nZVI. This material was then subjected to further characterization. The surface morphology of the synthesized material was studied using FESEM (Figure 2(a)) and TEM (Figure 2(b) and 2(c)). The FESEM image shows spherical and aggregated particles, while the TEM image shows polydispersed particles with spherical shapes. Particle size, determined using Image J software, was found to range between 20 and 72 nm, with a mean of 44 nm and a standard deviation of 15 nm. A 3D surface topographical pictograph was studied using AFM (Figure 2(e)), which confirmed the spherical shape of the particles. Analysis of AFM image (Figure 2(f)) in NOVA software shows most of the particles to be in the nano range, with the maximum number of particles in the size range of 50–70 nm. Rings and bright spots in the SAED image indicate the material to be crystalline.
Figure 2

Images of nZVI in (a) FESEM, (b, c) TEM, (d) SAED, and (e, f) AFM and particle size, respectively.

Figure 2

Images of nZVI in (a) FESEM, (b, c) TEM, (d) SAED, and (e, f) AFM and particle size, respectively.

Close modal
EDX showed Fe as the most prominent element (weight: 42%), followed by oxygen, sulfur, and potassium (weight: 37, 16, and 4%, respectively) (Figure 3(a)). The source of iron, oxygen, and sulfur can be FeSO4·7H2O, while the source of potassium can be leaf extract.
Figure 3

(a) EDX of synthesized nZVI, (b) FTIR of Sal leaf extract and nZVI before degradation, and (c) XRD of nZVI.

Figure 3

(a) EDX of synthesized nZVI, (b) FTIR of Sal leaf extract and nZVI before degradation, and (c) XRD of nZVI.

Close modal

FTIR was performed for Sal leaf extract and nZVI in the range of 500–4,000 cm−1 (Figure 3(b)). The broad peak at 3,323 cm−1 could be the stretching vibration of the sample's hydroxyl group of alcohol, carboxyl, phenol, or water, while the broad peak at 3,286 cm−1 could be due to O–H stretching in polyphenols. The peak marked C–O stretching due to primary alcohol at 1,045 and 1,230 cm−1, while C = O stretching due to ketones and alkyl group was marked by the peak at 1,654 cm−1. The presence of the C–H bond of alkyl group was confirmed by the peaks at 2,854 and 2,924 cm−1. Sal leaf contained alcohols and polyphenols, which were possibly responsible for the reduction of precursor to nZVI and acted as the capping agent for nZVI, preventing it from rapid oxidation (Rashtbari et al. 2020). The presence of C = O stretching as well as O–H stretching was marked by the peak at 1,658 and 3,317 cm−1. Polyphenol in the nZVI suggested its role as the capping agent, while the O–H group can play a role in the degradation of TC and CIP.

XRD of nZVI was performed in the 2θ range 5–85°. The characteristic peak of Fe0 at 44° confirmed the presence of nZVI (Qu et al. 2020). Various oxides of iron were present in the synthesized material, as confirmed by the XRD peaks. The UV–Visible spectra of nZVI indicated its peak at 215 and 264 nm (Supplementary Figure S2). Similar results were obtained by Pattanayak & Nayak (2013).

Optimization of reaction variables for best degradation

The antibiotic concentrations varied from 10 to 35 mg L−1, with a linear increase of 5 mg L−1. For both antibiotics, the degradation efficiency first increased with an increase in concentration and then gradually decreased. At 15 and 25 mg L−1 of TC and CIP, optimum degradation efficiencies of 80.8 and 77.6% were achieved, respectively. This can be attributed to the fact that an increase in antibiotic concentration increases the number of antibiotic molecules, thus increasing the number of molecules reaching the active sites (Gupta et al. 2020). Once the equilibrium is attained, active sites of nZVI get saturated. Due to the lack of sufficient active sites in nZVI, a further increase in antibiotic concentration reduces degradation efficiency (Gupta et al. 2020). With an increase in the concentration of antibiotics, many intermediates are generated, which can compete with antibiotics to react with active radicals. Moradi et al. (2020) reported a similar trend while degrading sulfamethoxazole using a graphene-based nanocomposite.

At 15 and 25 mg L−1 of TC and CIP, respectively, pH was varied from 2 to 12 for both antibiotics. pH has a significant role in forming active species and free radicals. A low degradation efficiency was seen at pH 2 and pH 10. Removal efficiency first increased with an increase in pH and then decreased. High removal efficiency was seen in the pH range 4–8 for both antibiotics. The maximum degradation efficiency of 82.3% for TC at pH 6 and 77.9% for CIP at pH 6 were achieved (Figures 4(b) and 5(b)). At low pH, hydrogen ions are present in the solution, which can react with free radicals such as OH. The free radicals, responsible for the degradation of antibiotics, get exhausted while reacting with hydrogen ions, reducing the degradation efficiency (Huang et al. 2020; Moradi et al. 2020). nZVI has a pHzpc around 8 (Tarekegn et al. 2021). TC, which has pKa values of 3.3, 7.7, and 9.7, can stay as cations, zwitterions, and anions (Li et al. 2021; Tang et al. 2021). Once the solution turns basic, at pH 10, TC, CIP, and nZVI attain a negative charge, leading to repulsion between them, which could reduce degradation efficiency (Gupta et al. 2020; Manea et al. 2021). Thus, further experiments were carried out at pH 6.
Figure 4

Process variable optimization of (a) initial concentration of TC, (b) pH of the TC solution, (c) dose of nZVI, (d) magnetic stirring rate of solution during degradation, and (e) degradation time for TC using a UV–Visible spectrophotometer.

Figure 4

Process variable optimization of (a) initial concentration of TC, (b) pH of the TC solution, (c) dose of nZVI, (d) magnetic stirring rate of solution during degradation, and (e) degradation time for TC using a UV–Visible spectrophotometer.

Close modal
Figure 5

Process variable optimization of (a) initial concentration of CIP, (b) pH of the CIP solution, (c) dose of nZVI, (d) magnetic stirring rate of solution during degradation, and (e) degradation time for degradation study of CIP using a UV–Visible spectrophotometer.

Figure 5

Process variable optimization of (a) initial concentration of CIP, (b) pH of the CIP solution, (c) dose of nZVI, (d) magnetic stirring rate of solution during degradation, and (e) degradation time for degradation study of CIP using a UV–Visible spectrophotometer.

Close modal

The dose of nZVI was varied from 0.0035 to 0.0210 g L−1, with a constant interval of 0.0035 g L−1 for TC and 0.0070 to 0.0245 g L−1, with a continuous interval of 0.0035 g L−1 for CIP. With the increase in the dose of nZVI, the removal efficiency of both antibiotics gradually increased to 83.8% for TC at 0.0140 g L−1 and 84.5% for CIP at 0.0175 g L−1 (Figures 4(c) and 5(c)). The efficiency then reduced slowly with a further increase in the dose of nZVI. This is because an increase in nZVI dose increases the number of active sites, generates more active radicals, and thus enhances the degradation rate in the initial phases. Ren et al. (2020) observed a similar trend in the degradation of TC by a graphene-based photocatalyst. However, further addition of nZVI led to a decrease in degradation efficiency, which might be due to the agglomeration of catalyst particles, or increased scattering of light and reduced light penetration, which ultimately hampers the photocatalysis reactions (Gupta et al. 2020). Gupta et al. (2020) studied the degradation of C17H18FN3O3 using an S–C3N4/ZnO catalyst at a dose of 0.5–2.5 g L−1. In his degradation study, the degradation efficiency increased continuously with an increase in dose from 0.5 to 1.5 g L−1 and then decreased steadily up to 2.5 g L−1. Moradi et al. (2020) degraded sulfamethoxazole and pharmaceutical wastewater using MgO/ZnO/Graphene ternary nanocomposite. With the increase in catalyst dosage, he observed that the active sites acted as recombination centers, self-consuming powerful oxidative radicals, and reducing adsorption of light due to the light scattering effect.

Stirring was performed using a magnetic stirrer from 100 to 500 rpm to maximize the degradation of TC and CIP. Stirring improves the contact between antibiotic molecules and nZVI as well as increasing the dissolved oxygen level, resulting in a decline in the electron–hole recombination and generation of superoxide radicals (Ahmadpour et al. 2020). Maximum degradation efficiency for TC was achieved at 400 rpm for TC and 100 rpm for CIP (Figures 4(d) and 5(d)). At these optimized conditions, photocatalytic degradation efficiency was monitored at an interval of every 10 min. Degradation efficiency increased gradually up to 70 min for both TC and CIP. Maximum efficiency of 88 and 84% was recorded for TC and CIP, respectively (Figure 4(e)).

The rate of photocatalytic degradation of TC and CIP and reaction kinetics was computed on the optimized concentration of 15 mg L−1 for TC and 25 mg L−1 for CIP, pH in the range of 4–6, nZVI dosing of 0.0140 g L−1 for TC and 0.0175 g L−1 for CIP, in magnetic stirring condition at a rate of 400 and 100 rpm for TC and CIP, respectively at 70 min. The kinetic study suggested that both reactions followed pseudo-second-order, as indicated by the respective regression correlation coefficient (R2) values. The curves were plotted based on Equations (2) and (3) (Jha & Chakraborty 2020). R2 was determined to be 0.86 and 0.93, respectively for the pseudo-first and the pseudo-second-order reaction of TC, and 0.56 and 0.89, respectively, for the pseudo-first and the pseudo-second-order reaction of CIP (Figure 6(a) and 6(b)).
(2)
(3)
where Co represents the original concentration of antibiotics (mg L−1), while C indicates the concentration (mg L−1) at time t, while k1 is the rate constant for the pseudo-first-order, and k2 is the rate constant for the pseudo-second-order. Values of k2 were calculated to be 0.2232 min−1 and 0.03959 min−1 for TC and CIP from the plot of (1/C) − (1/Co) versus t.
Figure 6

(a) Reaction kinetics for TC degradation, (b) reaction kinetics for CIP degradation, (c) mass spectrometry of pure TC, (d) mass spectrometry of degraded TC solution, (e) mass spectrometry of pure CIP, and (f) mass spectrometry of degraded CIP solution.

Figure 6

(a) Reaction kinetics for TC degradation, (b) reaction kinetics for CIP degradation, (c) mass spectrometry of pure TC, (d) mass spectrometry of degraded TC solution, (e) mass spectrometry of pure CIP, and (f) mass spectrometry of degraded CIP solution.

Close modal

Probable mechanism of degradation of TC and CIP

The probable mechanism of degradation of TC and CIP by nZVI is shown in Figure 7. The core of nZVI is made up of Fe0, while the external shell is made up of FeOOH (Yang et al. 2019). In an aqueous solution, Fe0 releases e, and gets converted to Fe2+ (Equation (4)). Fe2+ then reacts with oxygen in water to produce superoxide ion, which are highly reactive species and helps in the degradation of antibiotics (Equation (5)). Fe2+ also oxidizes to Fe3+ by reacting with hydrogen ion and oxygen in water, while producing hydrogen peroxide, a strong oxidizing agent (Equation (6)). This hydrogen peroxide breaks into hydroxide radical and hydroxide ion, by reacting with Fe2+ (Equation (7)). Superoxide ion, electrons, hydroxide radical, and hydroxide ion can degrade antibiotics (Fu et al. 2015). The absorbance peak of TC and CIP peak decreased over time, probably resulting from their degradation (Figures 4 and 5). Mass spectroscopy of TC and CIP also indicates their degradation. The sharp peak of TC can be seen at 445 (m/z) in Figure 6(c), which is completely absent in the degraded TC solution, while various new peaks indicate its breakdown into smaller intermediates in Figure 6(d). Similarly for CIP, a strong peak at 332 (m/z) can be seen for pure compound (Figure 6(e)), while it is entirely absent in the degraded CIP solution, while various new peaks indicate its breakdown to smaller compounds (Figure 6(f)).
(4)
(5)
(6)
(7)
(8)
Figure 7

Possible mechanism of degradation of TC and CIP using nZVI.

Figure 7

Possible mechanism of degradation of TC and CIP using nZVI.

Close modal

The degradation of TC could be brought about by attack of the reactive species on the dimethylamino group, leading to the removal of methyl group, followed by breaking of double bond and removal of the amide group. The other pathway is based on the removal of methyl, amide, and dimethylamino groups, followed by amino groups. Degradation is brought about by ring opening, oxidation and removal of carbonyl, ester, and hydroxyl groups, opening or relocation of double bonds, the opening of long carbon bonds, oxidation of phenolic hydroxyl, and benzyl groups (Ma et al. 2022).

Similarly, the degradation of CIP is initiated by the opening of the piperazine ring, or the removal of the F atom, carboxyl group, or cyclopropyl group. The C–N bond of the piperazine ring gets oxidized, opening the ring structure, while the removal of the F atom leads to defluorination, where the C–F bond is substituted by –OH. The double bond breaks to the single bond, and –OH enters the ring and breaks open the structure, leading to oxidation of the cyclopropyl group due to electrons and holes generated by light irradiation (Hu et al. 2020).

Toxicity assay of the degraded TC and CIP solution

Bacterial strains from river water were isolated after serial dilution (10−5) on nutrient agar plates. Two distinct colony-forming species, namely SRB1 and SRB2 (Figure 8(a) and 8(b)), were isolated as axenic strains based on the color of the colonies and found to be gram-positive (Figure 8(c1) and 8(c2)). The bacterial strains, when subjected to antibiotic solutions through cup assay, displayed a clear zone of inhibition for TC at 15 mg L−1 and CIP at 25 mg L−1. However, the zone of inhibition was absent for the degraded TC and CIP solutions and Millipore water (used as control). This result indicates that the degraded solution was non-toxic to the bacterial strains implying the efficient photocatalytic degradation capability of nZVI. This fact also bears the testimony that these antibiotic residues in the effluent can be discharged into natural water bodies, as it would pose no toxicity to the aquatic microorganisms (Figure 8(d1)–8(d4)).
Figure 8

(a) Bacterial strains on nutrient agar plates, (b) two distinct colony-forming species, namely SRB1 and SRB2; (c1 and c2) gram-stained microscopic images of SRB1 and SRB2, respectively; (d1 and d2) toxicity assay of the degraded TC solution by SRB1 and SRB2; and (d3 and d4) toxicity assay of the degraded CIP solution by SRB1 and SRB2.

Figure 8

(a) Bacterial strains on nutrient agar plates, (b) two distinct colony-forming species, namely SRB1 and SRB2; (c1 and c2) gram-stained microscopic images of SRB1 and SRB2, respectively; (d1 and d2) toxicity assay of the degraded TC solution by SRB1 and SRB2; and (d3 and d4) toxicity assay of the degraded CIP solution by SRB1 and SRB2.

Close modal

Treatment cost analysis

To synthesize a liter of nZVI solution, 500 ml of 0.01 M FeSO4·7H2O solution and 500 ml of H2O are required. A liter of solution produces 0.7 g of nZVI powder. The process uses electricity to shake the solution using an orbital shaker. This brings the cost of 0.7 g to Rs. 1.45 (0.0208 USD) and the rate to Rs. 2.07 (0.0297 USD) per gram. To treat the 10,000 L of TC and CIP, 140 and 175 g of nZVI are required, respectively. Electricity is required for UV light and magnetic stirring of the solution. The cost of treatment of 10,000 L of TC and CIP is ₹ 989.80 (14.158 USD) and ₹ 1062.30 (15.198 USD).

Item (for 10,000 L antibiotics)RateRequired quantity for TCCost for TCRequired quantity for CIPCost for CIP
nZVI 2.07 INR/0.0297 USD per grams 140 g 289.8 INR/ 4.158 USD 175 g 362.3 INR/5.198 USD 
Electricity 7 INR/0.1 USD/kWh 100 kWh 700 INR/10 USD 100 kWh 700 INR/10 USD 
  Total cost 989.8 INR/ 14.158 USD Total cost 1062.3 INR/15.198 USD 
Item (for 10,000 L antibiotics)RateRequired quantity for TCCost for TCRequired quantity for CIPCost for CIP
nZVI 2.07 INR/0.0297 USD per grams 140 g 289.8 INR/ 4.158 USD 175 g 362.3 INR/5.198 USD 
Electricity 7 INR/0.1 USD/kWh 100 kWh 700 INR/10 USD 100 kWh 700 INR/10 USD 
  Total cost 989.8 INR/ 14.158 USD Total cost 1062.3 INR/15.198 USD 
Item (1 L nZVI)RateRequired quantityCost
FeSO4·7H2750/kg 1.39 g 1.04 (0.015 USD) 
Sal leaf extract  50 g  
Electricity 7/kWh 0.0585 kWh 0.41 (0.0058 USD) 
Total cost of synthesis of 0.7 g nZVI 1.45 (0.0208 USD) 
Item (1 L nZVI)RateRequired quantityCost
FeSO4·7H2750/kg 1.39 g 1.04 (0.015 USD) 
Sal leaf extract  50 g  
Electricity 7/kWh 0.0585 kWh 0.41 (0.0058 USD) 
Total cost of synthesis of 0.7 g nZVI 1.45 (0.0208 USD) 

The study revealed the successful synthesis of nano zero-valent iron (nZVI) particles through the bio-reductive pathway of Fe precursor in the presence of Sal leaf extract. The spherical crystalline particles with an average size of 44 nm indicated the presence of polyphenol and alcohol functional groups, possibly acting as reducing agents to the precursor and capping agent of nZVI. These particles, when applied for photocatalysis of antibiotics TC and CIP under UV illumination, achieved a removal efficiency of 88 and 84% in optimal conditions of 15 and 25 mg L−1 antibiotic concentrations, 0.014 and 0.0175 g L−1 doses of nZVI, in the pH range 4–6 in 70 min, respectively. Kinetic studies revealed pseudo-second-order reactions. Active species (·O, ·OH, and H+) generated by nZVI in the presence of UV light might be attributed to playing a significant role in the degradation of the antibiotics. When further verified using mass spectrometry, the degradation of antibiotics was confirmed into smaller breakdown products. Microbial assay confirmed the non-toxic nature of the breakdown products, implying that effluent containing residual antibiotics can be treated by this process and discharged into natural water bodies. Furthermore, the cost analysis of the entire process proved economically feasible for application in a more extensive setup.

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

The authors declare there is no conflict.

Abbasnia
A.
,
Zarei
A.
,
Yeganeh
M.
,
Sobhi
H. R.
,
Gholami
M.
&
Esrafili
A.
2022
Removal of tetracycline antibiotics by adsorption and photocatalytic-degradation processes in aqueous solutions using metal organic frameworks (MOFs): a systematic review
.
Inorganic Chemistry Communications
145
,
109959
.
https://doi.org/10.1016/j.inoche.2022.109959
.
Ahmadpour
N.
,
Sayadi
M. H.
,
Sobhani
S.
&
Hajiani
M.
2020
A potential natural solar light active photocatalyst using magnetic ZnFe2O4@ TiO2/Cu nanocomposite as a high performance and recyclable platform for degradation of naproxen from aqueous solution
.
Journal of Cleaner Production
268
,
122023
.
https://doi.org/10.1016/j.jclepro.2020.122023
.
Al Omari
S.
,
Al Mir
H.
,
Wrayde
S.
,
Merhabi
S.
,
Dhaybi
I.
,
Jamal
S.
,
Chahine
M.
,
Bayaa
R.
,
Tourba
F.
,
Tantawi
H.
,
Al Atrouni
A.
,
El Omari
K.
,
Ayash
O.
,
Zeidan
N. S.
,
Mallat
H.
,
Dabboussi
F.
,
Hamze
M.
&
Osman
M.
2019
First Lebanese Antibiotic Awareness Week campaign: knowledge, attitudes and practices towards antibiotics
.
The Journal of Hospital Infection
101
(
4
),
475
479
.
https://doi.org/10.1016/j.jhin.2018.07.009
.
Ben
Y.
,
Fu
C.
,
Hu
M.
,
Liu
L.
,
Wong
M. H.
&
Zheng
C.
2019
Human health risk assessment of antibiotic resistance associated with antibiotic residues in the environment: a review
.
Environmental Research
169
,
483
493
.
https://doi.org/10.1016/j.envres.2018.11.040
.
Bhagat
C.
,
Kumar
M.
,
Tyagi
V. K.
&
Mohapatra
P. K.
2020
Proclivities for prevalence and treatment of antibiotics in the ambient water: a review
.
npj Clean Water
3
(
1
),
1
18
.
Che
H.
,
Che
G.
,
Jiang
E.
,
Liu
C.
,
Dong
H.
&
Li
C.
2018
A novel Z-scheme CdS/Bi3O4Cl heterostructure for photocatalytic degradation of antibiotics: mineralization activity, degradation pathways and mechanism insight
.
Journal of the Taiwan Institute of Chemical Engineers
91
,
224
234
.
https://doi.org/10.1016/j.jtice.2018.05.004
.
Ezzariai
A.
,
Hafidi
M.
,
Khadra
A.
,
Aemig
Q.
,
El Fels
L.
,
Barret
M.
,
Merlina
G.
,
Patureau
D.
&
Pinelli
E.
2018
Human and veterinary antibiotics during composting of sludge or manure: global perspectives on persistence, degradation, and resistance genes
.
Journal of Hazardous Materials
359
,
465
481
.
https://doi.org/10.1016/j.jhazmat.2018.07.092
.
Ezzatahmadi
N.
,
Ayoko
G. A.
,
Millar
G. J.
,
Speight
R.
,
Yan
C.
,
Li
J.
,
Li
S.
,
Zhu
J.
&
Xi
Y.
2017
Clay-supported nanoscale zero-valent iron composite materials for the remediation of contaminated aqueous solutions: a review
.
Chemical Engineering Journal
312
,
336
350
.
https://doi.org/10.1016/j.cej.2016.11.154
.
Fu
Y.
,
Peng
L.
,
Zeng
Q.
,
Yang
Y.
,
Song
H.
,
Shao
J.
,
Liu
S.
&
Gu
J.
2015
High efficient removal of tetracycline from solution by degradation and flocculation with nanoscale zerovalent iron
.
Chemical Engineering Journal
270
,
631
640
.
https://doi.org/10.1016/j.cej.2015.02.070
.
Guo
Y.
,
Huang
W.
,
Chen
B.
,
Zhao
Y.
,
Liu
D.
,
Sun
Y.
&
Gong
B.
2017
Removal of tetracycline from aqueous solution by MCM-41-zeolite A loaded nano zero valent iron: synthesis, characteristic, adsorption performance and mechanism
.
Journal of Hazardous Materials
339
,
22
32
.
https://doi.org/10.1016/j.jhazmat.2017.06.006
.
Gupta
B.
,
Gupta
A. K.
,
Tiwary
C. S.
&
Ghosal
P. S.
2020
A multivariate modeling and experimental realization of photocatalytic system of engineered S–C3N4/ZnO hybrid for ciprofloxacin removal: Influencing factors and degradation pathways
.
Environmental Research
196
,
110390
.
https://doi.org/https://doi.org/10.1016/j.envres.2020.110390.
Hou
X.
,
Shi
J.
,
Wang
N.
,
Wen
Z.
,
Sun
M.
,
Qu
J.
&
Hu
Q.
2020
Removal of antibiotic tetracycline by metal-organic framework MIL-101(Cr) loaded nano zero-valent iron
.
Journal of Molecular Liquids
313
,
113512
.
https://doi.org/10.1016/j.molliq.2020.113512
.
Hu
X.
,
Hu
X.
,
Peng
Q.
,
Zhou
L.
,
Tan
X.
,
Jiang
L.
,
Tang
C.
,
Wang
H.
,
Liu
S.
,
Wang
Y.
&
Ning
Z.
2020
Mechanisms underlying the photocatalytic degradation pathway of ciprofloxacin with heterogeneous TiO2
.
Chemical Engineering Journal
380
,
122366
.
https://doi.org/10.1016/j.cej.2019.122366
.
Huang
Y.
,
Nengzi
L.
,
Zhang
X.
,
Gou
J.
,
Gao
Y.
,
Zhu
G.
,
Cheng
Q.
&
Cheng
X.
2020
Catalytic degradation of ciprofloxacin by magnetic CuS/Fe2O3/Mn2O3 nanocomposite activated peroxymonosulfate: Influence factors, degradation pathways and reaction mechanism
.
Chemical Engineering Journal
388
,
124274
.
https://doi.org/https://doi.org/10.1016/j.cej.2020.124274
.
Le
T.-H.
,
Ng
C.
,
Tran
N. H.
,
Chen
H.
&
Gin
K. Y.-H.
2018
Removal of antibiotic residues, antibiotic resistant bacteria and antibiotic resistance genes in municipal wastewater by membrane bioreactor systems
.
Water Research
145
,
498
508
.
https://doi.org/10.1016/j.watres.2018.08.060
.
Leili
M.
,
Fazlzadeh
M.
&
Bhatnagar
A.
2018
Green synthesis of nano-zero-valent iron from Nettle and Thyme leaf extracts and their application for the removal of cephalexin antibiotic from aqueous solutions
.
Environmental Technology
39
(
9
),
1158
1172
.
https://doi.org/10.1080/09593330.2017.1323956
.
Li
X.
&
Liu
L.
2021
Recent advances in nanoscale zero-valent iron/oxidant system as a treatment for contaminated water and soil
.
Journal of Environmental Chemical Engineering
9
(
5
),
106276
.
https://doi.org/10.1016/j.jece.2021.106276
.
Li
Q.
,
Zhao
S.
&
Wang
Y.
2021
Mechanism of oxytetracycline removal by coconut shell biochar loaded with nano-zero-valent iron
.
International Journal of Environmental Research and Public Health
18
(
24
),
13107
.
https://doi.org/10.3390/ijerph182413107
.
Ma
Y.
,
Sun
M.
,
Mominou
N.
,
Zuo
N.
,
Li
S.
,
Jing
C.
&
Wang
L.
2022
Efficient degradation of tetracycline over Co/La oriented on g-C3N4 with the outstanding oxygen activation property under visible light
.
Journal of Environmental Chemical Engineering
10
(
3
),
107916
.
https://doi.org/10.1016/j.jece.2022.107916
.
Manea
Y. K.
,
Khan
A. M.
,
Wani
A. A.
,
Qashqoosh
M. T. A.
,
Shahadat
M.
&
Salem
M. A. S.
2021
Hydrothermally synthesized mesoporous CS-g-PA@TSM functional nanocomposite for efficient photocatalytic degradation of Ciprofloxacin and treatment of metal ions
.
Journal of Molecular Liquids
335
,
116144
.
https://doi.org/https://doi.org/10.1016/j.molliq.2021.116144.
Moradi
S.
,
Sobhgol
S. A.
,
Hayati
F.
,
Isari
A. A.
,
Kakavandi
B.
,
Bashardoust
P.
&
Anvaripour
B.
2020
Performance and reaction mechanism of MgO/ZnO/Graphene ternary nanocomposite in coupling with LED and ultrasound waves for the degradation of sulfamethoxazole and pharmaceutical wastewater
.
Separation and Purification Technology
251
,
117373
.
https://doi.org/https://doi.org/10.1016/j.seppur.2020.117373.
Nawaz
M.
,
Khan
A. A.
,
Hussain
A.
,
Jang
J.
,
Jung
H.-Y.
&
Lee
D. S.
2020
Reduced graphene oxide − TiO2/sodium alginate 3-dimensional structure aerogel for enhanced photocatalytic degradation of ibuprofen and sulfamethoxazole
.
Chemosphere
261
,
127702
.
https://doi.org/10.1016/j.chemosphere.2020.127702
.
Orimolade
B. O.
,
Zwane
B. N.
,
Koiki
B. A.
,
Tshwenya
L.
,
Peleyeju
G. M.
,
Mabuba
N.
,
Zhou
M.
&
Arotiba
O. A.
2020
Solar photoelectrocatalytic degradation of ciprofloxacin at a FTO/BiVO4/MnO2 anode: kinetics, intermediate products and degradation pathway studies
.
Journal of Environmental Chemical Engineering
8
(
1
),
103607
.
https://doi.org/10.1016/j.jece.2019.103607
.
Pattanayak
M.
&
Nayak
P. L.
2013
Green synthesis and characterization of zero valent iron nanoparticles from the leaf extract of Azadirachta indica (Neem)
.
World Journal of Nano Science & Technology
2
(
1
),
06
09
.
https://doi.org/10.5829/idosi.wjnst.2013.2.1.21132
.
Qu
G.
,
Chu
R.
,
Wang
H.
,
Wang
T.
,
Zhang
Z.
,
Qiang
H.
,
Liang
D.
&
Hu
S.
2020
Simultaneous removal of chromium(VI) and tetracycline hydrochloride from simulated wastewater by nanoscale zero-valent iron/copper-activated persulfate
.
Environmental Science and Pollution Research
27
(
32
),
40826
40836
.
https://doi.org/10.1007/s11356-020-10120-8
.
Rashtbari
Y.
,
Hazrati
S.
,
Azari
A.
,
Afshin
S.
,
Fazlzadeh
M.
&
Vosoughi
M.
2020
A novel, eco-friendly and green synthesis of PPAC-ZnO and PPAC-nZVI nanocomposite using pomegranate peel: cephalexin adsorption experiments, mechanisms, isotherms and kinetics
.
Advanced Powder Technology
31
(
4
),
1612
1623
.
https://doi.org/10.1016/j.apt.2020.02.001
.
Ren
F.
,
Zhu
W.
,
Zhao
J.
,
Liu
H.
,
Zhang
X.
,
Zhang
H.
,
Zhu
H.
,
Peng
Y.
&
Wang
B.
2020
Nitrogen-doped graphene oxide aerogel anchored with spinel CoFe2O4 nanoparticles for rapid degradation of tetracycline
.
Separation and Purification Technology
241
,
116690
.
https://doi.org/https://doi.org/10.1016/j.seppur.2020.116690.
Selvamani
P. S.
,
Vijaya
J. J.
,
Kennedy
L. J.
,
Mustafa
A.
,
Bououdina
M.
,
Sophia
P. J.
&
Ramalingam
R. J.
2021
Synergic effect of Cu2O/MoS2/rGO for the sonophotocatalytic degradation of tetracycline and ciprofloxacin antibiotics
.
Ceramics International
47
(
3
),
4226
4237
.
https://doi.org/https://doi.org/10.1016/j.ceramint.2020.09.301.
Stefaniuk
M.
,
Oleszczuk
P.
&
Ok
Y. S.
2016
Review on nano zerovalent iron (nZVI): from synthesis to environmental applications
.
Chemical Engineering Journal
287
,
618
632
.
https://doi.org/10.1016/j.cej.2015.11.046
.
Tan
X.
,
Liu
Y.
,
Zeng
G.
,
Wang
X.
,
Hu
X.
,
Gu
Y.
&
Yang
Z.
2015
Application of biochar for the removal of pollutants from aqueous solutions
.
Chemosphere
125
,
70
85
.
https://doi.org/10.1016/j.chemosphere.2014.12.058
.
Tang
X.
,
Huang
Y.
,
He
Q.
,
Wang
Y.
,
Zheng
H.
&
Hu
Y.
2021
Adsorption of tetracycline antibiotics by nitrilotriacetic acid modified magnetic chitosan-based microspheres from aqueous solutions
.
Environmental Technology & Innovation
24
,
101895
.
https://doi.org/https://doi.org/10.1016/j.eti.2021.101895.
Tarekegn
M. M.
,
Hiruy
A. M.
&
Dekebo
A. H.
2021
Nano zero valent iron (nZVI) particles for the removal of heavy metals (Cd2+, Cu2+ and Pb2+) from aqueous solutions
.
RSC Advances
11
(
30
),
18539
18551
.
https://doi.org/10.1039/D1RA01427G
.
Tran
M. L.
,
Nguyen
C. H.
,
Tran
T. T. V.
&
Juang
R.-S.
2020
One-pot synthesis of bimetallic Pt/nZVI nanocomposites for enhanced removal of oxytetracycline: roles of morphology changes and Pt catalysis
.
Journal of the Taiwan Institute of Chemical Engineers
111
,
130
140
.
https://doi.org/10.1016/j.jtice.2020.05.001
.
Winckler
C.
&
Grafe
A.
2001
Use of veterinary drugs in intensive animal production
.
Journal of Soils and Sediments
1
(
2
),
66
70
.
https://doi.org/10.1007/BF02987711
.
Yang
J.
,
Hou
B.
,
Wang
J.
,
Tian
B.
,
Bi
J.
,
Wang
N.
,
Li
X.
&
Huang
X.
2019
Nanomaterials for the removal of heavy metals from wastewater
.
Nanomaterials
9
(
3
),
424
.
https://doi.org/10.3390/nano9030424
.
Zhao
J.
,
Yang
X.
,
Liang
G.
,
Wang
Z.
,
Li
S.
,
Wang
Z.
&
Xie
X.
2020
Effective removal of two fluoroquinolone antibiotics by PEG-4000 stabilized nanoscale zero-valent iron supported onto zeolite (PZ-NZVI)
.
Science of the Total Environment
710
,
136289
.
https://doi.org/10.1016/j.scitotenv.2019.136289
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Supplementary data