1 Introduction

Over the past decades, there has been great interest emanating from many studies on the increasing prevalence of various pollutants in rivers, lakes, groundwater, and soil originating from a large variety of activities including chemical and petrochemical industries, pharmaceutical and veterinary industries, in addition to personal care products. In addition, the recent COVID-19 pandemic caused an increase in the global production of medicines for the purpose of addressing the health consequences of this disease including for products to diagnose, mitigate, or prevent COVID-19, and for other critically needed products to treat symptoms of COVID-19 or to provide supportive care to those with COVID-19. This, in turn, will certainly lead to the entry of some of different compounds into the aquatic environment, which is associated with risks for human health and pollution of ecosystems. Among many pharmaceutical pollutants in the environment, antibiotics are widely used in the fields of combating and in the treatment of human and veterinary bacterial diseases [1, 2]. For instance, one of the extensively employed antibiotics is ampicillin. Ampicillin is one of the β-lactam antibiotics [3]; it is used to treat a wide range of bacterial infections affecting humans and livestock, such as bladder, ear, respiratory, reproductive system, and gastrointestinal tract infections [4,5,6]. Through binding to peptidoglycan-synthesizing enzymes, ampicillin hinders the synthesis of the cell wall of bacteria.

In fact, the environmental contamination by antibiotics can lead to serious consequences for human health and aquatic biota depending upon the pollutant’s concentration and exposure time. Although the contamination from antibiotics is limited (normally in concentrations below the µg/L level), their risks lie in long-term exposure via drinking water or food to these compounds. Moreover, one of the important issues which has been identified for presence of this kind of pollutants in the environment is the development of the phenomenon of antibiotic-resistant bacterial diseases [7]. In addition, these compounds have the possibility of producing highly toxic or carcinogenic byproducts during the course of their degradation in the environment, or their interaction with other natural or anthropogenic pollutants [8,9,10]. The most important example of this is the interaction of pharmaceutical compounds with chlorine, or ozone added for disinfection purposes in drinking-water plants [11]. Therefore, there is an urgent need to address and mitigate the negative effects of the environmental pollution with antibiotics by applying an efficient method of treatment of wastewater containing antibiotics [12,13,14]. The focus is currently on devising simple, cost-effective, and efficient methods that ensure reducing these pollutants to the internationally permissible limits [13, 15, 16].

Recently, different techniques have been applied in pharmaceutical compounds for the purpose of removal of aqueous solutions including flocculation, flotation, photocatalysis, membranes, adsorption, and oxidation techniques [17, 18]. Among the multitude of treatments, the adsorption method has attracted substantial attention by several environmental researchers in the treatment of wastewater containing hazardous pollutants. Adsorption is universal and is the preferential treatment process for removing different types of elements because of its high efficiency, simplicity in design and operation, high-quality effluents, cost-effectiveness, and feasibility [19, 20]. Major mechanisms responsible for the removal of pollutants by the adsorption include chemical reactions, bonding processes, electrostatic processes, and ion exchange. From another side, the adsorption process is a sensitive process and its applicability and efficiency in the removal are highly dependent on the nature of the adsorbent used. Many adsorbents, such as commercial activated carbon (CAC), have been found to possess a high capacity for removing diverse types of pollutants from liquid wastes [21,22,23]. However, CAC is not a cheap adsorbent especially when it was used in a large-scale treatment process, and the regeneration process of this adsorbent is complex [24,25,26]. The best technique to overcome these disadvantages is the derivation of activated carbon (AC) from a low-cost plant or agricultural wastes, which is the main objective of this study, and which has been reported as a good alternative compared to expensive CACs [27]. Hence, algae, ferns, and other natural plants are one of the best carbon sources, and they are actually one of the cheapest raw materials as well. Among the natural plants, Azolla filiculoides (AF), which is a floating water fern, has evoked a great interest in the water treatment using the adsorption technique [28, 29]. This fern is grown in bulk in several locations in the world, like in the north of Iran, and is usually consumed as fodder by livestock, or used as fertilizer [30, 31].

The purpose of this work was to prepare AC from AF (named ACAF) and examine its adsorption capacity to remove ampicillin from polluted solutions. The chemical properties, surface, and structural properties were determined using powerful characterization techniques. In addition, the adsorption performance of ACAF for the removal of ampicillin was examined at different values of environmental parameters. The isotherm, kinetic, and thermodynamic capacities of ampicillin adsorption onto ACAF were also discussed. Ultimately, the regeneration potential of the ACAF was tested through five adsorption–desorption cycles.

2 Materials and methods

2.1 Chemicals

Ampicillin (chemical formula: C16H19N3O4S, molecular mass: 349.4 g/mole) was purchased from Sigma-Aldrich Co., USA. The stock solution of ampicillin (1000 mg/L) and other working solutions were prepared via dilution using double-distilled water. The other reagents/chemicals used in the current work were of analytical grade.

2.2 Adsorbent preparation

The ACAF was prepared from AF using both chemical and physical activation techniques. The chemical activation was accomplished by using ZnCl2 as a dehydration agent, whereas the physical activation was carried out by a heating process at high temperature. First, the required quantity of AF was obtained from an open area in Anzali wetland (Iran). Note that the unit cost of collecting 1 kg of fresh AF was < US$ 1. Next, the collected AF was dried, crushed to a powder, and finally sieved to uniform particles (0.45 mm). Subsequently, the AF powder was soaked in a ZnCl2 solution (28%) for 12 h. After that, this mixture was placed in a stainless steel reactor (placed in a temperature-controlled vertical furnace) under a high-purity N2 atmosphere at a flow rate of 300 cm3/min. The mixture was gradually heated to a temperature of 450 °C and kept at this temperature for 1 h. At this stage, the resultant solid particles were ACAF, which was subsequently subjected to acid treatment using hot HCl (0.5 M) in order to remove excess ZnCl2. The ACAF was separated from solution via filtration, washed with warm water, dried in a digital oven at 105 °C for 24 h. The percentage yield of synthesized ACAF can be determined using Eq. 1 [32].

$$Yield=\frac{{W}_{C}}{W}\times 100$$
(1)

where \({W}_{C}\) and \(W\) are the weights of obtained ACAF and dried AF, respectively.

2.3 Characterization of ACAF

The changes in the topographical characterization of the ACAF before and after ampicillin adsorption process were observed by FESEM images, which were taken using high-resolution scanning electron microscopy (S-4800F, HITACHI). N2 adsorption–desorption technique (at a temperature of =  − 196 °C) using a Micromeritics-TRISTAR-3000 surface area and porosity analyzer was utilized for exploration of the surface area, porosity, and total pore volume of ACAF. The elemental analysis was conducted by using Perkin Elmer elemental analyzer (Model EA2400II). FTIR was performed using a model device Thermo Nicollet AVATAR 5700 using the KBr pellet method in the range 400–4000 cm−1. The experimental steps of analysis of point of zero charge (pHpzc) are shown in the supplementary information file (Section S1).

2.4 Experiments and calculations

The adsorption experiments were conducted in a series of 200-mL Erlenmeyer flasks filled with 100 mL of ampicillin solution of predetermined concentration. The adsorbents added in each flask were stirred with an ampicillin solution at stirring speed of 150 rpm. The ACAF adsorption capacity for ampicillin was determined at different conditions: pH (3–11) which is regulated through dropwise addition of NaOH and/or HCl solution (0.1 M), initial ampicillin concentration (25–100 mg/L), adsorption time (0–120 min), ACAF dose (0.1–0.5 g/L), temperature (15–45 °C), and concentration of competitor NaNO3 ion (0–0.5 mol/L). The regeneration experiment of spent ACAF was performed in five consecutive adsorption–desorption cycles of ampicillin under the following experimental conditions: ACAF dose = 0.8 g/L, pH = 7, initial ampicillin concentration = 25 mg/L, adsorption time = 60 min, and temperature = 25 °C.

At a specific time of adsorption reaction, a 3-mL sample was taken from each solution and quickly centrifuged at 1509 × g for 10 min. Determination of the initial and remaining ampicillin concentration in solutions was done using HPLC (Kenaver HPLC device) with C18 column, and detection was measured at λ = 215 nm with a diode array UV–Vis detector after centrifugation of solution sample (10 mL) at 3600 rpm for 10 min. The dihydrogen phosphate (pH = 3.5)-acetonitrile (87.5:12.5, v/v) was the mobile phase employed, which delivered at a flow rate of 1.00 mL/min. The removal efficiency (\(\mathrm{\% removed}\)) and adsorption capacity (\({q}_{t})\) of ampicillin molecules onto ACAF were determined using Equations (S1) and (S2) shown in the supplementary information file (Section S2). For the kinetic and thermodynamic studies, the results of the experiments of the ACAF adsorption capacity at different ampicillin initial concentrations and temperatures were used, respectively. In addition, an experiment was performed to determine the relationship between the ACAF adsorption capacities toward ampicillin at a different temperature (15–45 °C), and the results of this experiment were used for the isotherm study. In addition, for selecting the best models, the determination correlation coefficient (R2) and normalized standard deviations Δq (%) (Equation S3: supplementary information file) values were employed in the isotherm and kinetic studies.

3 Results and discussion

3.1 Characterization analyses

It was found that the main elements that formed ACAF are carbon (51.2%), hydrogen (4.18%), nitrogen (0.73%), and oxygen (40.71%). The ash content in the ACAF sample was 2.84%. Using Eq. 1, the percentage yield of synthesized ACAF adsorbent from AF fern was found to be 44.7% at the cost of US$ 10/kg. Table 1 lists the results of the specific surface area, pore size, and other characteristics of ACAF.

Table 1 Surface and pore characteristics of ACAF

Figure 1 illustrates the FESEM images of the ACAF before and after ampicillin adsorption. From the FESEM image before adsorption, the ACAF can be described as having many pores as well as having many fragments scattered on its surface and inside the pores. In fact, these properties, which are highly favorable in the adsorption process, are characterized by many carbon adsorbents synthesized from agricultural materials, and this results from exposure of these materials to high temperatures during the synthesizing process [33]. After ampicillin adsorption, it can be seen that the surface characterization of used adsorbent was completely changed. All the previous detected pores disappeared and the surface became more flat. In addition, the fragments became larger. These changes can be reasoned to have occurred from the attachment and sorption of ampicillin molecules on the surface and pores of ACAF.

Fig. 1
figure 1

SEM images of ACAF before and after ampicillin adsorption

The N2 adsorption–desorption and pore size distribution analyses of ACAF are presented in Fig. 2. This figure was indicative of A type IV isotherm by the hysteresis loop at high P/P0 values. Moreover, the pore diameter range of 4–5 nm was determined for ACAF particles.

Fig. 2
figure 2

N2 adsorption–desorption isotherms and BJH pore size distribution curves

Nomination of bonds in the synthesized ACAF adsorbent (before and after ampicillin adsorption) was done using FTIR analysis (in the range of 400–4000 cm−1) (Fig. 3). In this figure, shifting and disappearance of some peaks and the appearance of some new peaks could be noticeable, which was representative of this fact that, in the adsorption of ampicillin, the functional groups on the surface of the ACAF are affective. Table 2 lists the details of the FTIR analysis employed for our studied adsorbent. From this table, it can be seen that it was the functional groups which were responsible for adsorption of the ampicillin molecules.

Fig. 3
figure 3

FTIR spectra of ACAF samples (temperature = 150 °C)

Table 2 Functional groups that were detected in the FTIR spectra of ACAF samples before and after ampicillin adsorption

3.2 Effect of ACAF dosage

By varying the ACAF dosage in the aqueous solution (0.1–1.5 g/L), the evaluation of effect of ACAF dose on the adsorption efficiency of ampicillin was reckoned (Fig. 4). As elucidated by this figure, an improvement in ampicillin removal from aqueous was detected as the ACAF mass was increased; however, for dose of ACAF > 0.8 g/L, no change in efficiency values was observed. Accordingly, an enhancement in ampicillin adsorption efficiency from 39.26 to 96.84% was noticed due to an increase in dosage from 0.1 to 0.8 g/L; this results in the improvement in the number of active sites for adsorption [34]. Moreover, as illustrated in Fig. 4, the qe can be affected by effect of ACAF mass; so that an increase in ACAF mass (0.1 to 1.5 g/L) leads to diminish it (98.15 to 16.26 mg/g). This may be explained according to the collision of adsorbent particles with each other and the formation of bulk; this reduces the active sites for adsorption of ampicillin [35].

Fig. 4
figure 4

Effect of ACAF dosage on efficiency of ampicillin adsorption (C0: 25 mg/L, temperature = 30 ± 2 ℃, pH: 7, time: 60 min)

3.3 Effect of pH and mechanism of adsorption

The variation of pH values has been detected to be associated with a remarkable effect on the adsorbent surface and ampicillin ionization in solution [36]. The results of studying this parameter are shown in Fig. 5. The values of pK1 (–COOH) and pK2 (–NH3+) of ampicillin have been detected to be 2.5 and 7.3, respectively. Thus, deprotonation of protonated ampicillin containing –COOH and –NH3+occurs in the pH ranges 2–4 and 6–8, respectively [37]. According to the results, removal percentage was developed with an increase in pH values from 3 to 7; however, after that, a gradual decrease in the removal efficiency up to a pH of 11 was detected. Moreover, at a pH of 7, the highest removal percentage of 96.84% was detected. Describing this event is done based on this fact that both ACAF and ampicillin have positive charges at pH < 2.5; this diminishes the adsorption of ampicillin onto ACAF surface [38]. However, deprotonation of ampicillin occurs at a pH of 7. This leads to formation of negative sites to interact with the positive surface of ACAF, which provides the highest adsorption of ampicillin in the mentioned pH value (pH = 7). In addition, the adsorption of ampicillin was declined at pH values higher than 7. The formation of electrostatic repulsion forces between carboxylate groups present in ampicillin molecules and the negative surface charges of the ACAF could describe the observed event in these pH values [39]. Moreover, it is expected that π-π interaction between a benzene ring of ampicillin and ACAF happens, which leads to the developing of adsorption capacity [40].

Fig. 5
figure 5

Effect of pH on the ampicillin adsorption onto ACAF (C0: 100 mg/L, temperature = 30 ± 2 ℃, time: 60 min, ACAF dosage: 0.8 g/L)

As shown in Figure S1, the point of zero charge (pHzpc), which can be useful for justifying adsorption at a low pH, was 7.6. ACAF in the solution pH lower than the mentioned value, that is, 7.6, have positive charges, while for the solution with a pH higher than 7.6, the adsorbent had negative charges [41]. Therefore, the presence of a negative charge in the surface of ACAF at pH values higher than 7.6 leads to a decline in the removal efficiency due to electrostatic repulsion between negatively charged ampicillin species and the negatively charged ACAF. Another reason for this event is the existence of a higher concentration of OH ions in the working solutions [42].

3.4 Effects of initial ampicillin concentration and kinetic study

The examination of the changes in ampicillin removal efficiency by varying the initial ampicillin concentration was done using 100 mL solution of initial concentrations of 25, 50, 100, and 200 mg/L and 0.8 mg of ACAF, and the results of this part are reported in Fig. 6. In this figure, rapid ampicillin adsorption is detected in the beginning and after that, some minor enhancement is observed until it reaches equilibrium. Since the active site on ACAF surface is more accessible in the start of process, adsorption of ampicillin speedily occurs at the initial contact time. As mentioned, there is a slow rate of adsorption after this stage. A description of this reason can be done based on the slow pore diffusion of the adsorbate molecule into the bulk of the adsorbent [43, 44]. Since the maximum adsorption rate was obtained at 60 min and the process reached equilibrium at this contact time, 60 min was considered equilibrium contact time for the next experiments. In addition, qe (adsorption capacity at equilibrium) was observed to be developed from 30.35 to 189.25 mg/g by increasing the initial ampicillin concentrations (from 25 to 200 mg/L). By increasing the initial concentration of adsorbate, the driving force of the concentration gradient is developed, which may be the reason for the enhancement of qe [45].

Fig. 6
figure 6

Effect of contact time on the adsorption capacity of ampicillin (pH = 7, ACAF dosage 0.8 g/L, and temperature = 30 ± 2 ℃)

Analyzing the kinetic data (Fig. 6) was done by pseudo-first-order (PFO) (Equation S4) and pseudo-second-order (PSO) (Equation S5) models [46, 47]. More information about the theoretical background used in this study is presented in the supplementary material file (Section S2). The kinetic parameters (e.g., correlation coefficients, R2, and normalized standard deviations, Δq (%)), which are determined using the linearized kinetic plots of PFO and PSO (Fig. 7), are documented in Table 3. Through calculating R2, and Δq (%), and the agreement between the calculated qe cal and the experimental qe exp values, the best kinetic model for describing the obtained data is determined. As determined, for PSO, higher R2 values were obtained compared to the PFO model. In addition, the values of Δq (%) for PSO were found to be much lower compared to the values for PFO model. In addition, it was found that the closeness of values of the qe cal and the qe exp for PSO model was greater than the PFO model. The above results are indicative of more appropriateness of the PSO model for enlightening the behavior of the studied process. Furthermore, increasing the initial ampicillin concentration and surface loading has led to a decrease in the values of K2. With an increase in initial concentration of studied pollutant and surface loadings, diffusion efficiency is declined and the competition of ampicillin molecules for fixed reaction sites is raised. This results in lower K2 values.

Fig. 7
figure 7

Pseudo-second-order kinetic plots for adsorption of ampicillin on ACAF

Table 3 The results of kinetic model studies related to the ampicillin adsorption onto ACAF

Owing to the inability of the kinetic models mentioned above for defining a diffusion mechanism, the kinetic data were fitted by intra-particle diffusion model (IPD) model (Equation Section S5: supplementary material). When the mechanism of adsorption process obeys the IPD, drawing the qt vs. t1/2 provides a straight line; the slope and intercept of the plot are used for the K and I, respectively. As can be seen in Fig. 8, IPD plots were not linear over time. This highlights the existence of more than one mechanism affecting the ampicillin adsorption for ACA. Furthermore, the plots do not pass through the origin. According to the foregoing, the presence of IPD is confirmed, but in addition to this, there are other mechanisms that act as rate-controlling steps. In Table 3, values obtained for K and I have been reported, According to this table, an increase in initial concentration is associated with enhancing K values, which is explained by an increase in the driving force that occurs due to initial concentration and adsorption through mesopores and micropores [47]. In addition, I values were observed to be greater than 0, and increasing the initial ampicillin concentration has led to the enhancement of its values. Considering this fact, boundary layer diffusion might be the rate-limiting step in the adsorption process, not intra-particle diffusion [48].

Fig. 8
figure 8

Intra-particle diffusion kinetic plots for adsorption of ampicillin on ACAF

3.5 Equilibrium isotherms

In this study, the equilibrium data was modeled with four isotherm models, including Freundlich, Langmuir, Dubinin-Radushkevitch, and Temkin (Section Section S3: supplementary information). The results of each model’s parameters are listed in Table 4. It can be seen that the determined range of RL value was 0.043 to 0.177 at temperatures of 15–45 °C. This is demonstrative of the favorable adsorption of ampicillin onto ACAF. In addition, a diminution in RL values was detected when the temperature increased from 15 to 45 °C. This reveals that higher temperatures are significantly favorable for the adsorption of ampicillin. According to the results, the highest ampicillin adsorption capacities onto ACAF at different temperatures were 89.92, 96.74, 105.2, and 114.3 mg/g, respectively. These values were comparable to values obtained for some other adsorbents. The details are provided in Table 4.

Table 4 Isotherm parameters for adsorption of ampicillin onto ACAF at various temperatures

According to the values of n, the adsorption process is linear when n = 1, the chemical adsorption occurs when n < 1, and the adsorption is a physical process when n > 1 [49]. The intercept and slope of the plot of lnqe against lnCe (not shown) are employed for determining constants of KF and n. In this study, according to the results of the Freundlich isotherm, the values of n were observed to be less than 1 for all studied temperatures. This specifies the physisorption nature of the process. In addition, an increase in values of the Freundlich constant (KF) at higher temperatures is representative of the favorability of higher temperatures for the adsorption process. Furthermore, since, in the present study, the 1/n value observed at temperatures studied was lower than unity, and adsorption of ampicillin onto ACAF was favorable.

The calculation of constants of K and qm of D-R model is done through employing the slope and intercept of the plot, which was obtained from drawing lnqe against ε2 (not shown). E values obtained were 3.54, 4.92, 6.11, and 7.35 kJ/mol, at 15, 25, 35, and 45 °C, respectively. The mentioned values were observed to be less than 8 kJ/mol under studied experimental conditions (Table 2); this highlights the physisorption nature for adsorption of ampicillin onto ACAF.

Calculating b and KT of the Temkin model is done based on the slope and intercept of the line which is obtained by drawing qe against lnCe (not shown). This is representative of improving the heat of adsorption of ampicillin onto the surface of ACAF when temperature (from 15 to 45 °C) increases. It also agrees with the endothermic and thermodynamic results found for sorption. Moreover, since values of b were observed to be lower than 80 kJ/mol, a physical adsorption process occurred. Comparison of correlation coefficient (R2) achieved for studied isotherm models (Table 4), aptness of the Langmuir model for portraying the adsorptive behavior was confirmed since R2 of this model was greater than the other studied models. Moreover, on the basis of the results of the best fit model, homogenous surface of the ACAF was predicted for the adsorption of ampicillin molecules from aqueous solution.

In Table 5, the highest adsorption capacity of ACAF for adsorption of ampicillin has been compared with the others. According to the results of this comparison, the adsorption capacity of ACAF is high compared to most of the reported adsorbents, which indicates the suitability of this adsorbent for use in the adsorption processes.

Table 5 Comparison of adsorption capacities of various adsorbents for ampicillin

3.6 Adsorption thermodynamics

Considering the results related to this part of the study, the values for ΔGo were observed to be negative; so that − 2.94, − 3.84, − 5.73, and − 8.69 kJ/mol were calculated at 15, 25, 35, and 45 °C, respectively. On the basis of those negative values, the adsorption process was considered to be spontaneous [50, 51]; nevertheless, increasing the temperature has provided the enhancement in absolute values of ΔGo. It indicates that the adsorption is more feasible when the temperature is high. Moreover, according to values of ΔGo (within the ranges of − 20 and 0 kJ/mol), the adsorption mechanism was identified to be mainly physical [52]. In addition, the value of ΔHo obtained was 19.82 kJ/mol, which was a positive value. Positive values for this parameter are demonstrative of an endothermic nature for adsorption processes. In addition to clarification of the above-mentioned nature, the type of adsorption (either physical or chemical) can be detected based on the magnitude of ΔHo [53]. The values obtained for ΔHo (from 2.1 to 20.9 kJ/mol) highlight that fact that studied adsorption process is of the physisorption type. The positive value obtained for ΔSo (0.188 kJ/mol. K) represents the presence of the attraction of ACAF for ampicillin. It is also indicative of enhancing randomness at the solid-solution interface in the adsorption process [54].

3.7 Effect of ionic strength

Evaluating the effect of ionic strength on the adsorption of ampicillin was done through conducting experiments with NaNO3 (the electrolyte in the range of 0 to 0.5 mol/L). The possible competition which may occur between Na+ and NO3 ions and ampicillin adsorption for adsorption sites or formation of general effect of bionic atmosphere or saline effect can affect the adsorption of the pollutant onto studied adsorbent. Nevertheless, our observation provided in Fig. 9 is indicative of the immaterial effect of ionic strength on ampicillin adsorption. On the basis of this, a lack of competition between added ions and ampicillin for adsorption and a lack of influence of electrostatic shielding effect on adsorption capacity sites can be concluded [55].

Fig. 9
figure 9

Effect of ionic strength on the equilibrium sorption capacity of ampicillin by ACAF (time = 60 min, C0 = 25 mg/L, pH = 7, ACAF mass = 0.8 g/L, and temperature = 25 °C)

3.8 Reusability of ACAF

Regeneration studies which are conducted for examining the reusability of adsorbents are essential to comprehend whether the studied adsorbent is suitable for practical applications [56]; to understand the mentioned factor for adsorbent used in this study, and the adsorption–desorption process of ampicillin was considered for five cycles (Fig. 10). According to the figure, in each cycle, compared to the former cycle, a decrease in the adsorption capacity of ACAF could be detected. In cycle no. 1, the adsorption capacity of ACAF for 25 mg/L ampicillin was observed to be 30.35 mg/g (%97.12), while it reached 29.11 mg/g (93.17) in cycle no. 5. This low decrease is representative of the potential regeneration and reusability of ACAF in the adsorption treatment systems.

Fig. 10
figure 10

Effect of regeneration on the equilibrium sorption capacity of ampicillin by ACAF (time = 60 min, C0 = 25 mg/L, pH = 7, ACAF mass = 0.8 g/L, temperature = 25 °C)

4 Conclusion

According to the results of the present study, the BET surface area of chemically activated ACAF was high and equal to 716.4 m2/g. The SEM analysis employed was indicative of the rough and porous surface of the synthesized adsorbent; it confirms the good properties of the adsorbent properties and potential for employment in the adsorption processes. Additionally, using the Langmuir, Freundlich, Temkin, and D-R isotherm models, the experimental data were analyzed, and the better applicability of the Langmuir isotherm for describing the data was ratified. Moreover, based on correlation coefficients values (R2) obtained for kinetic models, PSO was more effective in explaining the data. Optimum values of the contact time and pH for offering highest adsorption efficiency were 60 and 7 min, respectively. The highest adsorption capacity of the studied adsorbent was identified to be about 114.3 mg/g (45 °C). This is representative of a high potential of it for ampicillin removal. Furthermore, the spontaneous and endothermic nature of ampicillin adsorption onto ACAF was proven based on thermodynamic studies. The above-mentioned results present ACAF as a promising adsorbent for removing residual ampicillin from polluted water.