Potential use of dry powder of Vossia cuspidata (Roxb.) Griff. rhizomes and leaves in methylene blue dye remediation

News

HomeHome / News / Potential use of dry powder of Vossia cuspidata (Roxb.) Griff. rhizomes and leaves in methylene blue dye remediation

Jul 17, 2023

Potential use of dry powder of Vossia cuspidata (Roxb.) Griff. rhizomes and leaves in methylene blue dye remediation

Scientific Reports volume 13, Article number: 11073 (2023) Cite this article 342 Accesses 4 Altmetric Metrics details Phytoremediation is a promising, cost-effective, and eco-friendly process for

Scientific Reports volume 13, Article number: 11073 (2023) Cite this article

342 Accesses

4 Altmetric

Metrics details

Phytoremediation is a promising, cost-effective, and eco-friendly process for wastewater treatment. Herein, the dry biomasses of Vossia cuspidata (Roxb.) Griff. leaves (PL) and rhizomes including aerial stems (PR) were used to effectively remediate methylene blue (MB) dyes. Interestingly, the adsorption uptake and removal efficiency of MB by PR were higher than those of PL; exceeding 97 and 91% in 35 and 25 min for 0.1 and 0.4 g/L MB, respectively. The MB diffusion within the PL and PR was insignificant and the adsorption kinetics was principally controlled by the surface MB–adsorbent interaction, as consistently approved by the pseudo-second order kinetic model. In addition, the adsorption increased rapidly with the plant dosage with high dependence on the initial MB concentration. Moreover, the impact of shaking speed on the adsorption was minor but temperature played a critical role where the highest efficiencies were recorded at 30 and 40 °C on PL (91.9%) and PR (93.3%), respectively. The best removal efficiencies were attained with PR at pH 6, but with PL at pH 8. The Temkin isotherm could perfectly simulate the experimental data (R2 > 0.97); suggesting a linear decrease of the adsorption heat of MB with the plant coverage.

According to United Nations World Water Development Report in 2020, around 4 billion people are subjected to severe physical water scarcity for at least 1 month in a year1,2. This number is expected to increase rapidly with the population growth to reach up to 10.2 billion people in 20503. In another estimate by the Food and Agriculture Organization (FAO) of the United Nations, 3.2 billion people live in agricultural areas with high to very high water shortages or scarcity, of whom 1.2 billion people—roughly one-sixth of the world’s population—live in severely water-constrained agricultural areas4. Hence, various plans should be proposed to remediate wastewater, particularly from sewage plants. Environmental pollution by dyestuff is a global issue that needs more attention for a healthy environment. Pollution with dyes causes hazardous impacts for all living organisms and depending on the ecosystem, they may further affect photosynthesis and food chains5. Typically, dyes are extremely toxic and carcinogenic, together with persisting natural degradations6. Unfortunately, these pollutants are widely used in considerable amounts in several textiles, cosmetics, food, pharmaceutical, paper, and leather industrial activities. The world’s annual production of dyes is more than 700 thousand tons. About 2% of them are directly discharged from manufacturing plants into effluents, and roughly 10% of dyes are lost throughout the dyeing industries process7,8,9,10. The discharge of these dyes in the ecosystem causes undesirable environmental loads11,12,13,14.

Methylene blue (MB, also known as methylthioninium chloride) is one of the most common substances that is widely used in dyeing industries. The presence of this dye even in low concentrations causes undesirable coloration of the aquatic ecosystem. Removal of (MB) using some natural raw materials such as rice husk, hair, and guava seeds was studied by several groups15,16. However, the management of this pollutant is still a growing issue, necessitating the development of new technologies, appealing to the ecological valorization concerns17. Among the recent methods for the decolorization of dye-containing effluents, the removal via adsorption onto specific adsorbents appeared highly promising from a considerable environmental perspective18. Despite the wide variety of sorbents in the literature, biomass derived from agricultural waste or plant residues appears highly attractive for the adsorption of emerging inorganic heavy metals and organic dyes contaminants19. Generally, biomass is cost-effective, environmentally benign and renewable biosorbent of high contents of carbon and cellulose, and diverse functionality20. This diversity is accustomed to biomass due to the availability of multiple surface functional –OH, –C=O and –COOH groups that furnish a high adsorption tendency to a large scale of adsorbates21,22. The role of activation of biosorbents and their applications to remediate wastewater from heavy metals and dyes were recently reviewed2,19.

Generally, in phytoremediation where biomass from plant residues is used to remediate contaminated industrial streams, the same principles and mechanistic steps of adsorption were applied. This involved a cascade of simultaneous events transporting the contaminant molecules from the bulk solution to the interface by migration, diffusion, and/or convection before getting adsorbed at the solid/H2O interface23. A structural modification of the adsorbate at the solid/H2O interface or within the pores of the adsorbent for better packing with lateral interactions of adsorbate molecules is possible23. In natural ecosystems, adsorption competition between molecules of different natures and molecular weights should be additionally considered24. Finally, the surface interaction (whether physically or chemically) of the adsorbate at the adsorbent surface and the possible desorption should, moreover, be addressed. Any of these steps may determine the kinetics of the adsorption process, and accordingly, several kinetics models and adsorption isotherms were developed25,26,27,28.

Numerous studies have been conducted on the use of plants and their biomass in the cleanup of contaminated environments29,30,31,32,33,34. More than 400 plants including many hydrophytes have been assigned as prospective phytoremediators35,36,37,38. For example, Neolamarckia cadamba leaves succeeded to remediate MB dye and to decrease significantly the total dissolved solids (~ 27%), conductivity (~ 21%), hardness (~ 89%) and chemical oxygen demand (~ 66%) of the sewage belts of Yamuna River in Allahabad, Uttar Pradesh, India39. Alternatively, the aquatic plant, Azolla pinnata as a biosorbent prospered to impart ~ 85% decolorization of MB dye (25 mg/L) from contaminated wastewater streams within 48 h40.

Vossia cuspidata (Roxb.) Griff. (family Poaceae) is a long-lived macrophyte grass that is widely grown in Southeast Asia and Tropical Africa41 The natural vegetation of the River Nile in Egypt has drastically decreased because of its recent invasion of the aquatic environment42. In Egypt, V. cuspidata was investigated for its ability to phytostabilize heavy metals from polluted water canal43,44,45,46. They reported that this species can accumulate significant levels of Cd, Cu, Mn, Pb, Ni, Co, and Zn ions in their roots. However, to our knowledge, there are no data available about the use of dry biomass of V. cuspidata in the phytoremediation of industrial MB dyes. Therefore, the objective of this study is to investigate the potential use of the dry biomass of V. cuspidata in the phytoremediation of textile MB dyes. The removal of MB dye was employed for marking the adsorption efficiency and capacity of the biomass of the V. cuspidata plant leaves and stems including the underground rhizomes and aerial stems.

The plant materials were collected from the natural habitats of V. cuspidata along the Nile River, Cairo, Egypt. Plant samples were transported to the laboratory and divided into rhizomes, aerial stems, and leaves. We refer to the V. cuspidata plant’s rhizomes and stems as PR, and plant leaves as PL. After separation, all organs were washed by running tap water several times to clean the plant from debris and dust, then washed with distilled water. Each plant organ's component was first air-dried, followed by oven drying at 60 °C for 72 h to achieve a constant mass, and finally ground to a fine powder in an agate mortar and sieved through a 0.2 mm mesh sieve. The dry biomass of PL and PR was used as an adsorbent for MB dye’s remediation. The functional groups of the adsorbent were characterized by Fourier Transform Infrared spectroscopy (FTIR, Model Bomen/MB102, ABB Company, Switzerland). The surface morphology of the V. cuspidata was evaluated by the field-emission scanning electron microscopy (FE-SEM, Quanta FEG250).

To test the ability of dry biomass of the selected organs of V. cuspidata to remediate the MB dye, several tests were conducted including the effect of contact time, initial adsorbate concentration and adsorbent mass, temperature, pH, and shaking speed. During the adsorption studies, the removal efficiency of MB onto the PL and PR was calculated using Eq. (1)47:

where \({C}_{e}\) (mg/L) is the final (equilibrium) MB dye (adsorbate) concentration and \({C}_{0}\) (mg/L) is the initial MB dye concentration in the bulk solution. All adsorption measurements were repeated three times before testing the curve fitting, calculating the standard error, and performing the statistical analysis.

1 g dry biomass of PL or PR was mixed with 100 mL of 0.1, 0.2, 0.3, and 0.4 g/L MB solution. Each solution was represented by three replicates to avoid a decrease in the pollutant’s volume. The solutions were kept at room temperature and natural pH. Supernatant samples of 10 mL were taken every 5 min and centrifuged then the remaining dye was analyzed at 664 nm using a spectrophotometer (PerkinElmer Lamda 25 UV/VIS spectrophotometer). The MB dye absorption at the same wavelength was calibrated.

To elucidate the adsorption kinetics, the amount (mg) of MB dye adsorbed by the dry plant biomass at a given time (\({q}_{t}\)) and at equilibrium (\({q}_{e}\)) were calculated according to Eqs. (2) and (3), respectively48:

where \({C}_{t}\) (mg/L) is the concentration of MB dye (adsorbate) at a given time in the bulk solution of a volume \(V\) (L). The mass of PL or PR biomass (adsorbent) is \(W\) (g).

The kinetics data of MB adsorption on the PL or PR biomass were tested for fitting in the pseudo-first-order (PFO), pseudo-second-order (PSO), and Weber and Morris (WM) intraparticle diffusion models48,49,50. According to the PFO rate model Eq. (4) and its linearized form Eq. (5) that was developed by Lagergren, \({q}_{t}\) increases exponentially with time, as following49,51:

where \({k}_{1}\) (min−1) is the first-order rate constant.

On the other hand, the PSO model proposes the following empirical linear formula Eq. (6) for \({q}_{t}\) vs. time52,53:

where and \({k}_{2}\) (mg/g/min) is the second order rate constant

The initial rate of adsorption (\(h\)) can be assessed according to the PSO model as follows Eq. (7)52,53;

Alternatively, the WM model suggested a linear variation of \({q}_{t}\) with the square root of time according to Eq. (8)48,50:

where \({k}_{3}\)(mg/g/min0.5) and \(C\) (mg/g) stand, respectively, for the rate and correlation constants of the intraparticle diffusion model.

Five different adsorbate concentrations (0.025, 0.05, 0.1, 0.2, and 0.4 g/L) of MB were prepared. In three separate reactors, 100 mL of each concentration of MB dye was mixed with one of these three loadings (1, 2, and 3 g) of PL and PR. The solutions were kept at room temperature and natural pH (≈ 7) to the desired contact time and then analyzed.

Six different loadings (0.5, 1, 1.5, 2, 2.5, and 3 g) of PL and PR dry biomass were mixed with 100 mL of 0.1, 0.2, 0.3, and 0.4 g/L MB solution, and solutions were kept at room temperature and natural pH (≈ 7). Supernatant samples of 10 mL were taken for analysis at the optimal contact time for each plant material.

Flasks containing 100 mL of 0.1 g/L MB solution were placed in a water bath to adjust their temperature before mixing with 1, 2, or 3 g of each plant material. The solutions were then incubated at seven different temperatures (20, 25, 30, 35, 40, 45, and 50 °C) and natural pH (≈ 7) in a water bath. At the optimal contact time, 10 mL of supernatant was taken for analysis.

100 mL of 0.1 g/L MB solution was treated with 1, 2, and 3 g of each plant material at five different pH (hydrogen ion) concentrations (6, 6.5, 7, 7.5, and 8, respectively). The tested solutions were adjusted to the desired pH value before mixing with the adsorbent using 0.1 M NaOH and 0.1 M HCl solutions. The mixtures were left to the equilibrium contact time before analysis.

Four adsorption models; namely, the Freundlich, Langmuir, Temkin, and Dubinin–Radushkevich (D–R) isotherms were employed. The Freundlich model which is concerned with the adsorption onto heterogeneous surfaces assumed the following relationship Eq. (9) between \({q}_{e}\) and \({C}_{e}\)54,55,56:

where \({k}_{F}\) is the adsorption capacity (L/mg) and n is an empirical constant inferring the relative distribution of energy and heterogeneity at the adsorbate surface54. For 1/n values between 0.1 < 1/n < 1, the adsorption process is favorable54.

On the other hand, the Langmuir isotherm that fitted perfectly for the adsorption of a monolayer of adsorbate onto the surface of the adsorbent assumed the uniformity of the adsorbent that ensures the equivalency of adsorption sites57. The adsorption of the adsorbate onto a finite number of identical active sites of the adsorbent with the same adsorption mechanism is also assumed. The interaction of the adsorbed molecules together, regardless of the surface coverage, is also forbidden58,59,60. Mathematically, the Langmuir adsorption isotherm model is represented by Eq. (10):

where \({q}_{m}\) (mg/g) is the maximum adsorption capacity and \({K}_{L}\) (L/mg) is the Langmuir equilibrium constant that is related to the energy of adsorption61.

However, the Temkin model is expressed in the following formula Eq. (11); accounting for the possible indirect adsorbate/adsorbent interactions48,56,62:

where \(b\) (J/mol) and \(A\) (L/g) are the Temkin constants, R is the gas constant (8.31 J/mol/K) and T (K) is the absolute temperature.

Finally, the D–R isotherm that realizes the surface's non-uniformity in accommodating different adsorption modes and fits perfectly for the pore-filling mechanism is expressed in the following Eqs. (12)–(14)56,63,64:

where \(\varepsilon\) is the Polanyi potential and \(\beta\) (mol2/kJ2) is the D–R constant or activity coefficient that can directly be converted to the average adsorption energy, \(E\), (kJ/mol).

100 mL of 0.1 g/L MB solution was mixed with (1, 2, and 3 g) of each plant material using five different shaking speeds (100, 150, 200, 250, and 300 rpm, respectively). The solutions were kept on the shakers at room temperature and natural pH (≈7) to the optimal contact time for each plant material, then moved to analysis.

Ethics approval in this study, plant materials were collected and used according to the national regulations. The utilization of these species for experimental purposes does not require any special permit. All methods in our study comply with relevant institutional, national, and international guidelines and legislation. The plant was identified by the last author according to Boulos31 and a voucher id (8105) for the plant was deposited at the herbarium of Helwan University.

Previous chemical analysis verified and estimated the contents of Na, K, N, Mg, P, Ca, ether, crude fibres, proteins, ash, and carbohydrates with seasonable variations in the aboveground leaves and stem, and in the belowground roots and rhizome of Vossia cuspidata65. Figure 1A,B show the FTIR of the PL and PR biomasses before “pristine” and after getting loaded with the MB dye. As seen in Fig. 1, the raw plant materials owned major peaks at 3417, 2924, 1646, and 1050 cm−1 for PL (Fig. 1A), and at 3424, 2925, 1639, and 1054 cm−1 for PR (Fig. 1B). Other minor peaks appeared at 1728, 1646, 1518, 1428, 1381, 1250, 1158, 902, 598, 467, and 429 cm−1 for PL (Fig. 1A), and at 1721, 1513, 1426, 1254, 1157, and 605 cm−1 for PR (Fig. 1B). Loading of MB onto the PL and PR biomasses resulted in obvious shifts in the peaks positions (Fig. 1A,B), which can correlate certainly to modulations of the vibrational energy structure of PL and PR due to their interaction with the MB dye.

FTIR spectra of V. cuspidata leaves, PL (A) and rhizome, PR (B) powder before and after MB loading.

The morphology of the V. cuspidata biomass before and after loading with MB was inspected. Figure 2A–D depicts the surface FE-SEM images of the PL (Fig. 2A,C) and PR (Fig. 2B,D) before and after soaking in MB-contaminated solution. As obviously seen in Fig. 2A,C, the PL displayed intensive microcavities at the surface with porous texture which is favorable for pore-filling adsorption mechanism. On the other hand, the PR biomass appeared as agglomerated tubular bundles. After loading with MB (Fig. 2C,D), the PL and PR surface appeared partially-coated with white debris.

FE-SEM images of the V. cuspidata leaves (A,C) and rhizome (B,D) before (A,B) and after (C,D) loading with MB-contaminated solution.

The experimental data for the MB adsorption by the PL and PR dry biomass were fitted in the PFO Eqs. (4) and (5), PSO Eq. (6), and WM Eq. (8) kinetics models (Fig. 3A–D, Tables 1, 2). The regression coefficient, R2, was assigned as a fitting index to evaluate the suitability of the different models to the adsorption data. It was revealed that the adsorption of MB onto both PL and PR powders was very rapid as \({q}_{t}\) increased semi-exponentially over time, jumping to saturation within the first few minutes (ca. 5 min) of adsorption (Fig. 3A). The equilibrium adsorption uptake, \({q}_{e}\), of PR (14.50 mg/g) was higher than that of PL (10.54 mg/g). After 5 min of contact with MB, 1 g of PR powder showed 89 and 85% removal when treated with 0.1 g/L and 0.4 g/L MB, respectively. On the other hand, the PL powder amounted to removal efficiencies of 91 and 64% when treated with the same concentrations, respectively. Increasing the contact time showed a slower removal rate. The removal efficiencies of MB onto PR powder extended to 97 and 91% at 35 and 25 min, respectively, for the same abovementioned solutions. However, 94 and 76% removal were recorded at 40 min for PL powder when treated with the same concentrations, respectively. While seemed adequate for the PFO kinetics model (Fig. 3A), the adsorption data could not be fulfilled perfectly in its linear form Eq. (5). This was verified for the adsorption data of MB onto PL and PR (Fig. 3B). On the other hand, the PSO kinetics model (Fig. 3C) appeared more appropriate for fitting the adsorption data of MB both onto PL and PR (R2 ≈ 1). The WM intra-particle diffusion model (Fig. 3D) did not perform well like the PSO model but was better than the PFO model with R2 = 0.79. This diminishes the significance of MB diffusion within the PL and PR and dedicates the adsorption kinetics principally to the surface MB–adsorbent interaction. Interestingly, the PSO model predicted that the amounts of MB that would adhere to the PL and PR at equilibrium would be around 10.63 and 15.02 mg/g, respectively when the initial concentration of MB was 0.1 g/L, which was exactly in line with the corresponding experimental data, \({q}_{e,exp}\), (10.63 mg/g for PL and 15.19 mg/g for PR). The slope and intercept of Fig. 3C were used to determine the rate constants, which were found to be 0.47 and 0.07 mg/g/min (Table 2). Consequently, the adsorption rates, \(h\), of 52.66 and 15.78 mg/g/min were assigned for the adsorption of MB onto PL and PR, respectively.

(A) Effect of contact time on the adsorption capacity of 0.1 g/L MB (natural pH) onto 1 g of PL (blue circles) and PR (red squares). The linear fitting of the experimental data was done by using the PFO (B), PSO (C), and WM (D) kinetic models.

The impact of the plant dosage on the adsorption uptake and efficiency is shown in Fig. 4. It is apparent that the MB adsorption increased rapidly with the plant dosage, because of increasing the adsorption sites available for the MB removal. Interestingly, 0.5 g PR powder induced ca. 76 and 73% removal efficiencies when treated with 0.1 g/L and 0.4 g/L of MB-containing solution, respectively. On the other hand, the same mass of PL powder recorded 75 and 70% removal efficiencies for the same solutions, respectively. Adsorption efficiencies up to 92.8 and 92.5% could be verified for the adsorption of MB onto the PL and PR, respectively, at higher doses of plant powder. Nonetheless, the influence of the plant mass on the adsorption capacity and efficiency of MB was minor within the range of 1 g/100 mL (10 g/L) to 2 g/100 mL (20 g/L) of PL and PR biomass, respectively. These results led us to believe that 10 g/L of PL and PR was the optimal dosage while also attempting to limit the variables.

Effect of plant dry biomass on the adsorption capacity (blue color) and removal efficiency (red color) of 0.1 g/L MB (natural pH) onto PL (circles) and PR (squares).

The investigation continued to analyze the role of the initial adsorbate concentration where several samples of MB with concentrations from 0.025 to 0.400 g/L were tested while maintaining the optimum adsorbent dosage of 10 g/L. From an industrial perspective, this analysis is highly important due to the presence of MB in variable concentrations in the industrial streams. In general, the removal tendency of plant materials was higher when treated with diluted MB solution. This ability decreased by increasing adsorbate concentration, but the actual amount of pollutant adsorbed per unit mass of adsorbent increased with increased adsorbate concentrations in the test solution. As shown in Fig. 5 a sample with 0.025 g/L MB showed removal efficiencies of 94 and 92% when treated with 1 g of PR and PL powder, respectively, while the same concentration amounted to 98% removal when treated with 3 g of both plant materials. On the other hand, a sample with 4 g/L MB attained removal efficiencies up to 89 and 87% when treated with 1 g and 96 and 98% when treated with 3 g of PR and PL powders, respectively.

Effect of adsorbate (MB) initial concentration on the adsorption capacity (blue color) and removal efficiency (red color) onto 10 g/L PL (circles) and PR (squares) at natural pH.

Figure 5 shows a linear increase of the equilibrium adsorption uptake of MB with the MB concentration up to 0.2 g/L with a minor distinction for PR over PL. This linear relationship continued up to 0.4 g/L MB but with lower slopes. The MB removal efficiency, on the other hand, depended largely on the initial MB concentration, and the highest efficiencies (up to 97 and 95% for PL and PR, respectively) were reported for intermediate concentrations (from 0.05 to 0.2 g/L MB). It is worth mentioning that the increase in the MB concentration to 0.4 g/L did not associate with a parallel and equivalent increase in the adsorption uptake, and hence, the removal efficiency decreased remarkably.

Four adsorption models—Freundlich, Langmuir, Temkin, and D-R isotherms—were used to simulate the interaction of MB dye at the surface of the PL and PR biomass and to determine their maximum adsorption capacities for the removal of MB dye.

Figure 6A shows the \({q}_{e}\) and \({C}_{e}\) correlation according to the Freundlich model Eq. (9) and evaluates the \({k}_{F}\), n and 1/n constants for the adsorption of MB dye onto the PL and PR biomass at room temperature. The R2 values of both plots in Fig. 6A were below 0.85, suggesting poor fittings of experimental data according to the Freundlich isotherm (Table 3). The values of \({k}_{F}\), n and 1/n were 1.70, 1.53, and 0.65, respectively, for PL and were 1.99, 1.30, and 0.77, respectively, for PR biomass. It is observed that 1/n value was low (< 1) while \({k}_{F}\) value for the adsorption of MB onto the PR was high.

Fitting the experimental adsorption data of MB according to the (A) Freundlich, (B) Langmuir, (C) Temkin, and (D) D–R adsorption isotherms using 10 g/L PL (blue circles) and PR (red squares) at natural pH.

On the other hand, to examine the Langmuir model’s applicability to the experimental data, Fig. 6B was plotted for \(\frac{{C}_{e}}{{q}_{e}}\) vs. \({C}_{e}\) (Eq. 10). It is interesting to note that the adsorption of MB onto the PL and PR was associated with regression factors (R2) higher than those found for the Freundlich model (see Table 3). The maximum monolayer adsorption capacities, \({q}_{m}\), for the MB adsorption onto PL and PR were 35.7 and 212.9 mg/g, while \({K}_{L}\) were 0.0329 and 0.0279 L/mg, respectively. According to the Langmuir model, the adsorption feasibility for different adsorbate concentrations is assessed by the separation factor (RL) which is a dimensionless parameter as seen in Eq. (15)57.

For \({R}_{L}>1\), the adsorption is unfavorable, for \({R}_{L}=1\), the adsorption is linear, for \({0<R}_{L}<1\), the adsorption is favorable, and for \({R}_{L}=0\), the adsorption is irreversible. Applying these criteria to the experimental data of MB adsorption onto the PL and PR using an initial MB dye concentration, C0, of 100 mg/L indicates that these adsorption processes are favorable.

The Temkin model Eq. (11) was next applied to test the \({q}_{e}\) vs. \({C}_{e}\) correlation as shown in Fig. 6C. It is revealed that the best correlation for the experimental data was obtained with this model, as indicated by the R2 (> 0.97) value (Table 3). For the adsorption of MB dye onto the PL and PR, the Temkin constants \(b\) were computed to be 333.18 and 279.52 J/mol while the constants \(A\) were 0.461 and 0.587 L/g, respectively (Fig. 6C).

Finally, the experimental data were tested for fitting according to the D − R isotherm (Eqs. 12–14). A simple plot of \(ln{q}_{e}\) vs. \({\varepsilon }^{2}\) (Fig. 6D) could predict \(\beta\) and \({q}_{m}\) from the slope and intercept, respectively (Table 3). A substitution in Eq. (14) could evaluate \(E\). Regression factors higher than 0.92 were obtained for modeling the adsorption data of MB dye onto the PL and PR according to this model. For the adsorption of MB dye onto the PL and PR, \({q}_{m}\) were calculated to be 19.78 and 24.67 mg/g while \(E\) were 0.31 and 0.38 kJ/mol, respectively (Fig. 6D). The results demonstrated that the model anticipated a higher \({q}_{m}\) for PR than PL for the removal of MB dye, which is in line with the experimental data. Additionally, based on the low E values, it determined that physisorption mechanisms were responsible for the removal of MB from both the PL and the PR dry biomass.

In the present study, the impact of shaking speed showed a minor effect on the adsorption of pollutants on PR and PL powders. When the shaking speed was increased from 100 to 300 rpm, 1 g of PR powder achieved up to 96% removal efficiency, and 3 g of the same plant material achieved up to 98% efficiency. Besides, the PL powder showed about 92 and 98% removal efficiency under the same conditions, respectively. The results on the effect of shaking speed between 100 and 300 RPM on the adsorption capacity and removal efficiency of 0.1 g/L MB using 10 g/L are shown in (Fig. 7). The optimization of the adsorption process at 150 RPM with an adsorption capacity of 9.34 mg/g and efficiency of 93.4% in PL was obtained. On the other hand, 200 RPM was optimized with an adsorption capacity of 9.68 mg/g and an efficiency of 96.8% for PR. It remains to visualize the adsorption proceeding in a reversible (adsorption/desorption) mode. Accelerated shaking rates may inspire different impacts on the rates of both the adsorption and desorption processes. Mechanically, high shaking rates can overcome the physical attraction forces between the sorbate and sorbent, enhancing the return of MB dye back to the solution and decreasing the adsorption uptake. That is essentially what was seen at shaking rates higher than 150 and 200 RPM in PL and PR, respectively. However, the higher tolerance of the MB-PR bonding for dissociation (desorption) early at 150 RPM was likely caused by the higher adsorption energy of MB onto PR (0.38 kJ/mol) than that of PL (0.31 kJ/mol).

Effect of shaking speed on the adsorption capacity (blue color) and removal efficiency (red color) of 0.1 g/L MB using 10 g/L PL (circles) and PR (squares) at natural pH.

Figure 8 showed that the equilibrium adsorption uptake of MB dye onto PL increased slightly from 20 to 30 °C then became approximately constant up to 40 °C, and then steadily decreased to 50 °C. The trend of MB dye adsorption onto PR was a little different as the equilibrium adsorption uptake decreased directly from 20 to 33 °C and increased again at 40 °C, and then continued decreasing up to 50 °C. The removal efficiencies did not change largely to the same extent as the adsorption uptake and the highest efficiencies were recorded at 30 and 40 °C on PL (91.9%) and PR (93.3%), respectively. The minor variation in the heat of adsorption of MB onto PL and PR, along with the geometrical differences between PL and PR, might be the reason for the observed behavior.

Effect of temperature on the adsorption capacity (blue color) and removal efficiency (red color) of 0.1 g/L MB using 10 g/L PL (circles) and PR (squares) at natural pH.

The impact of pH (6–8) on the adsorption process of MB dye was quite difficult (Fig. 9). This remained true for both PL and PR, showing that this pH range—close to neutrality—almost results in uptakes and efficiencies that were extremely similar. The removal efficiencies (94% using 1.0 g and 98% using 3.0 g of PR powder) were enhanced in the slightly acidic medium (pH 6). On the other hand, the PL powder showed the best removal efficiencies (93% and 98% using the same weights, respectively) in the slightly basic medium (pH 8).

Effect of pH on the adsorption capacity (blue color) and removal efficiency (red color) of 0.1 g/L MB using 10 g/L PL (circles) and PR (squares) at natural pH.

We are the first to report the functional group analysis of the PL and PR biomass of Vossia cuspidata. The major broad peaks at 3417 and 3424 cm−1 for PL and PR, respectively, belong mostly to the –NH/–OH overlapped stretching vibration, as commonly assigned in similar plants66,67,68. The other major peaks at 2924, 1646, and 1050 cm−1 for PL (Fig. 1A), and at 2925, 1639, and 1054 cm−1 for PR (Fig. 1B) can likely refer to the (C–H) asymmetric stretching, overlapped C=N stretching and N–H bending, and C–O–C and C–N overlapped stretching, respectively68,69,70,71. On the other hand, the minor absorption peaks of PL (Fig. 1A) were assigned to the C=O stretching of aldehyde (1728 cm−1), C=C stretching of imine/oxime (1646 cm−1), N–O stretching (1518 cm−1), C=C side ring stretching (1428 cm−1), –CH2 or –CH3 stretching (1381 cm−1), –C–N stretching (1250 cm−1), –C–O stretching of tertiary alcohol (1158 cm−1), C=C bending (902 cm−1), C–H bending (598 cm−1), Si–O–Si stretching (467 cm−1), and Si–O stretching (429 cm−1), respectively68,72,73,74. On the other hand, the minor peaks of PR biomass (Fig. 1B) were assigned similarly to the C=O stretching of aldehyde (1721 cm−1), N–O stretching (1513 cm−1), C=C side ring stretching (1426 cm−1), –C–N stretching (1254 cm−1), –C–O stretching of tertiary alcohol (1157 cm−1), C–H bending (605 cm−1)68,72. The loading of MB dye onto the PL biomass did not inspire great shifts in the peak positions, in contrast to the case of PR biomass. It is worth mentioning that the FTIR spectrum of the MB dye itself depicted very similar absorptions to most of those of unloaded PL and PR68. Further chemical analysis is required to understand the nature of the interaction between MB dye and PL and PR biomasses.

The surface morphology (Fig. 2) of the V. cuspidata biomass (PL and PR) appeared porous and bundled with dispersed microcavities, which is a favorable structure for the pore-filling adsorption mechanism. The parenchyma and aerenchyma as well as xylem vessels in both plant materials showed an irregular porous morphology at the plant surface. The white debris that were appeared after loading the PL and PR surface with MB (Fig. 2C,D) can likely return to MB.

In this study, the appropriateness of the PSO kinetic model was confirmed for fitting the adsorption of MB dye onto the dry powder of V. cuspidata. The same model was previously recommended for the adsorption of organic dyes by plant dry biomasses18,75,76. The equilibrium adsorption uptakes, \({q}_{e}\), of PR and PL were 14.50 and 10.54 mg/g. While removal efficiencies of 97 and 94% for MB were obtained onto the PR and PL powders (1 g/100 mL) after 35 and 40 min, respectively. These values stand among the highest reported values for MB from aqueous contaminated streams (Table 4)77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95.

The rapid adsorption rate of any chemical may be due to the presence of a substantial number of unoccupied adsorption sites existing on the adsorbent surface96. The presence of MB in contact with the powder of V. cuspidata rhizome and leaves causes a sudden accumulation of dye on the surface of the adsorbent. With time, these sites become occupied by MB which may produce a repulsive force between the molecules of dye on the adsorbent surface as reported by Mahmoodi et al.97. At higher dye concentrations, the number of dye molecules competing for the available sites on the surface of the adsorbent is large which may explain the faster equilibrium time. This rapid adsorption time is recommended and highly desirable in the application of this technology. When employing activated carbon made from Ficus carica wood and modified oak residue, the equilibrium contact time of MB was reached in 90 min and 120 min, respectively98,99. It was observed that the equilibrium times for MB removal onto V. cuspidata were shorter, which indicates better adsorbent materials.

Studying the adsorbent dose used in the adsorption process is a vital factor, as it determines the equilibrium conditions between the adsorbent and pollutants in the reaction system that can be used to evaluate the ecological cost-benefits during the treatment of polluted habitats100. Increasing the adsorbent dose causes an increase in the contact surface area with much more active sites for binding with MB. Many factors may contribute to the higher removal capacity, including solute availability, electrostatic interactions, interference between binding sites, and lower mixing at high biomass densities101. The findings of the present study agreed with that reported by Esmaeili and Foroutan102.

At constant adsorbent dose, V. cuspidata powders showed different removal rates when treated with different dye concentrations. This phenomenon is very much related to diffusion constraints. If the MB concentration is low, the ratio of MB to the number of available active adsorption sites is small, which encounters the mass transfer of MB from the aqueous to the PL and PR surface to yield low adsorption uptakes. On the other hand, higher initial MB concentrations may overcome the diffusion limitations and support higher MB fluxes to the plant surface, increasing the adsorption uptakes58,59,60. Concerning constant biomass weight, increasing dye concentration (up to 4 g/L) resulted in a slight decrease in removal capacity, which indicates the high ability of the plant biomass for adsorption wherever the position of adsorbent from the source point of pollution. Regarding the adsorption process, the quantity of adsorbed dyes increased as the initial dye concentration increased. The increase in the concentration gradient with increasing dye initial concentration could explain this phenomenon103.

On the other hand, the Freundlich adsorption isotherm model had low (< 1) 1/n values but high \({k}_{F}\) values for the adsorption of MB onto the PL and PR. Small numerical values for 1/n (< 1) not only confirm the favorable adsorption of MB onto the PL and PR but also indicate the nature of physical bonding between MB and the adsorbent54. The higher \({k}_{F}\) for the adsorption of MB onto the PR indicated a corresponding higher adsorption capacity. Likewise, a higher n value for the MB adsorption onto PL suggested a more favorable adsorption than onto PR. The Temkin model showed the best correlation for the experimental data compared to other tested models. Consequently, we can suggest the Temkin model for the adsorption of the MB dye onto the PL and PR. In a previous investigation, the same model was suggested (R2 = 0.95) for the adsorption of MB onto Miswak (Salvadora persica) leaves in an alkaline solution with an adsorption capacity of 200 mg/g104. Al-Ghouti and Al-Absi investigated the cationic methylene blue dye's adsorption and thermodynamic effects on cellulosic olive stone biomass18. They found that the green olive stones obtained the highest R2 value with the Temkin isotherm model, whereas the black olive stones showed the highest R2 value with the Langmuir isotherm model.

The same results were reported by Amin et al. for the batch adsorption of crystal violet and methylene blue dyes from an aqueous solution using Eucalyptus camdulensis biochar76. The Temkin isotherm was nearly a perfect fit to the adsorption data of the MB dye, with an R2 value close to unity (1.0). This significant correlation shows that the MB dye was adsorbed in a heterogeneous manner with a homogeneous distribution of binding energies.

The nature of adsorbent–adsorbate interaction can be predicted from data obtained from the investigated adsorption isotherms in this study. The Temkin constant \(b\) that is related to the adsorption energy was 333.18 and 279.52 J/mol for the adsorption of MB dye onto the PL and PR. The average adsorption energy, \(E\), for the adsorption of MB dye onto the PL and PR was estimated from the D–R isotherm as 0.31 and 0.38 kJ/mol, respectively, which agreed to a great extent with the Temkin constant. These findings came consistent with the values of 1/n (0.65 and 0.77 for the adsorption of MB onto PL and PR biomass, respectively) that were computed by the Freundlich adsorption isotherm model to indicate the nature of physical bonding between MB and the adsorbent54. One of the common physical bonding is the dipole–dipole interaction which is likely to prevail due to the existence of positive centers at the sulfur atom of MB and positive centers at the oxygen atoms of the hydroxyl groups that enrich the plant surface (as evidenced from IR spectra). The schematic in Fig. 10 illustrates this sort of attractions. More chemical analysis and materials characterizations are required to specifically assign the sort of the prevalent physical interaction.

Plausible adsorption mechanism for the removal of MB by V. cuspidata biomass. The green dotted lines represent the physical dipole–dipole interactions between MB and the plant.

To expedite the mass transfer of MB to the sorbent surface, the influence of shaking rate, a hydrodynamic parameter, was then examined. Noubactep et al. investigated previously the impact of shaking rate between 0 and 300 min−1 on the adsorption of MB onto scrap iron (Fe0), granular activated carbon (GAC), and deep-sea manganese nodules (MnO2)24. While non-shaken experiments lasted for up to 50 days, shaken experiments were executed in less than a day. Their results suggested 50 min−1 as the optimum shaking rate for the facilitated mass transfer and the highest MB discoloration with no damage or dissolution of the sorbent material. Shaking is an essential factor that affects the bombardments between adsorbate and active sites on adsorbent. Physically, there are two types of mass transfer resistances: external and internal. The external one is due to the external diffusion between solution and adsorbent particles, while the internal one is due to the diffusion within the particles. Changing the shaking speed causes not much change in removal efficiency75. This is an indicator of the negligible effect of external mass transfer resistance (external diffusion) at these shaking speeds.

One of the most important parameters influencing the rates of adsorption and desorption of metal ions and organic dyes is the operating temperature. High temperatures kinetically speed up both, facilitating attaining the equilibrium quickly at high equilibrium constants. The mass transfer of the adsorbate from/to bulk solution is also enhanced at elevated temperatures. In addition, the adsorption/desorption processes follow the equilibrium constraints, and raising the temperature would favor one direction at the expense of the other based on Le Chatelier’s principle. Thermodynamically, adsorption processes are always exothermic, which necessitates boosting the desorption process at elevated temperatures, decreasing the adsorption uptake. It is also known that by increasing the temperature, the viscosity of the adsorbate solution decreased, and the movement of dye particles increased. The slight change in the “high” adsorption rate within the wide range of tested temperatures (20–50 °C), indicates the reduced economic cost of the adsorption process requirements103. Accordingly, V. cuspidata adsorbent biomass can remove considerable amounts of dyes at a variety of temperatures which represents an effective adsorptive material.

The hydrogen ion (pH) concentration of any solution has a considerable effect on the adsorption process. The pH affects the ionization of the dye molecule as well as the charge of functional groups on the adsorbent materials105. Among the studied range of hydrogen ion concentration, our results in Fig. 9 showed that the adsorptive capacity of the rhizome powder was higher in the slightly acidic medium (pH 6), while leaf powder showed higher capacity in the slightly basic medium (pH 8). The difference in the favorable adsorption pH values between rhizome and leaves powders of V. cuspidata is likely due to the difference in chemical composition and active site charges of each adsorbent material. In an acidic medium, the positively charged adsorbent sites are increased and become more than the negatively charged sites. On the contrary, in the basic medium, the numbers of negatively charged sites are more than the positive ones106. The normal pH of natural environments varies from slightly acidic to slightly alkaline media, which provides ideal adsorption conditions for V. cuspidata powders107.

The dry biomass of V. cuspidata rhizomes including aerial stems and leaves could be used as efficient materials in the phytoremediation of MB dye from industrial sources that contaminates water bodies. A removal efficiency up to 97% was obtained in 35 min for 0.1 g/L MB. Increasing the plant dosage increased the adsorption uptake with high dependence on the initial MB concentration. The experimental data could perfectly be fitted in the Temkin isotherm model where the energy parameters suggested the physical interaction between the plant biomass and MB dye. Further research should be done on the optimal settings and packing characteristics for the plant powders according to industrial purposes.

All data generated during this study are included in this published article.

Water, U. Water and Climate Change. The United Nations World Water Development Report (UNESCO, 2020).

Google Scholar

Tee, G. T., Gok, X. Y. & Yong, W. F. Adsorption of pollutants in wastewater via biosorbents, nanoparticles and magnetic biosorbents: A review. Environ. Res. 212, 113248. https://doi.org/10.1016/j.envres.2022.113248 (2022).

Article CAS PubMed Google Scholar

Boretti, A. & Rosa, L. Reassessing the projections of the world water development report. NPJ Clean Water 2, 15 (2019).

Article Google Scholar

FAO. The State of Food and Agriculture 2020. Overcoming Water Challenges in Agriculture (FAO, 2020).

Google Scholar

Mittal, A., Mittal, J., Malviya, A. & Gupta, V. K. Adsorptive removal of hazardous anionic dye “Congo red” from wastewater using waste materials and recovery by desorption. J. Colloid Interface Sci. 340, 16–26. https://doi.org/10.1016/j.jcis.2009.08.019 (2009).

Article CAS PubMed Google Scholar

Gupta, S., Sundarrajan, M. & Rao, K. V. K. Tumor promotion by metanil yellow and malachite green during rat hepatocarcinogenesis is associated with dysregulated expression of cell cycle regulatory proteins. Teratog. Carcinog. Mutagen. 23, 301–312. https://doi.org/10.1002/tcm.10056 (2003).

Article CAS Google Scholar

Saravanan, R., Thirumal, E., Gupta, V. K., Narayanan, V. & Stephen, A. The photocatalytic activity of ZnO prepared by simple thermal decomposition method at various temperatures. J. Mol. Liq. 177, 394–401. https://doi.org/10.1016/j.molliq.2012.10.018 (2013).

Article CAS Google Scholar

Gupta, V. K., Jain, R. & Varshney, S. Removal of Reactofix golden yellow 3 RFN from aqueous solution using wheat husk—An agricultural waste. J. Hazard. Mater. 142, 443–448. https://doi.org/10.1016/j.jhazmat.2006.08.048 (2007).

Article CAS PubMed Google Scholar

Gupta, V. K., Rastogi, A. & Nayak, A. Adsorption studies on the removal of hexavalent chromium from aqueous solution using a low cost fertilizer industry waste material. J. Colloid Interface Sci. 342, 135–141. https://doi.org/10.1016/j.jcis.2009.09.065 (2010).

Article CAS PubMed Google Scholar

Mittal, A., Mittal, J., Malviya, A., Kaur, D. & Gupta, V. K. Decoloration treatment of a hazardous triarylmethane dye, Light Green SF (Yellowish) by waste material adsorbents. J. Colloid Interface Sci. 342, 518–527. https://doi.org/10.1016/j.jcis.2009.10.046 (2010).

Article CAS PubMed Google Scholar

Ayed, L., Mahdhi, A., Cheref, A. & Bakhrouf, A. Decolorization and degradation of azo dye Methyl Red by an isolated Sphingomonas paucimobilis: Biotoxicity and metabolites characterization. Desalination 274, 272–277. https://doi.org/10.1016/j.desal.2011.02.024 (2011).

Article CAS Google Scholar

Vinoda, B., Vinuth, M., Bodke, Y. & Manjanna, J. Photocatalytic degradation of toxic methyl red dye using silica nanoparticles synthesized from rice husk ash. J. Environ. Anal. Toxicol. 5, 2161 (2015).

Google Scholar

Ahmad, M. A., Ahmad, N. & Bello, O. S. Modified durian seed as adsorbent for the removal of methyl red dye from aqueous solutions. Appl. Water Sci. 5, 407–423. https://doi.org/10.1007/s13201-014-0208-4 (2015).

Article CAS Google Scholar

Dadfarnia, S., HajiShabani, A. M., Moradi, S. E. & Emami, S. Methyl red removal from water by iron based metal-organic frameworks loaded onto iron oxide nanoparticle adsorbent. Appl. Surf. Sci. 330, 85–93. https://doi.org/10.1016/j.apsusc.2014.12.196 (2015).

Article CAS Google Scholar

Rahman, I. A. & Saad, B. Utilization of guava seeds as a source of activated carbon for removal of methylene blue from aqueous solution. Malaysian J. Chem. 5, 8 (2003).

Google Scholar

McKay, G., Porter, J. F. & Prasad, G. R. The removal of dye colours from aqueous solutions by adsorption on low-cost materials. Water Air Soil Pollut. 114, 423–438. https://doi.org/10.1023/a:1005197308228 (1999).

Article CAS Google Scholar

Elçi, H. Utilisation of crushed floor and wall tile wastes as aggregate in concrete production. J. Clean. Prod. 112, 742–752. https://doi.org/10.1016/j.jclepro.2015.07.003 (2016).

Article Google Scholar

Al-Ghouti, M. A. & Al-Absi, R. S. Mechanistic understanding of the adsorption and thermodynamic aspects of cationic methylene blue dye onto cellulosic olive stones biomass from wastewater. Sci. Rep. 10, 15928. https://doi.org/10.1038/s41598-020-72996-3 (2020).

Article CAS PubMed PubMed Central Google Scholar

Chong, Z. T., Soh, L. S. & Yong, W. F. Valorization of agriculture wastes as biosorbents for adsorption of emerging pollutants: Modification, remediation and industry application. Results Eng. 17, 100960. https://doi.org/10.1016/j.rineng.2023.100960 (2023).

Article CAS Google Scholar

Li, H. et al. A cost-effective porous carbon derived from pomelo peel for the removal of methyl orange from aqueous solution. Colloids Surf. A Physicochem. Eng. Asp. 489, 191–199 (2016).

Article CAS Google Scholar

Redha, A. A. Removal of heavy metals from aqueous media by biosorption. Arab. J. Basic Appl. Sci. 27, 183–193 (2020).

Article Google Scholar

Din, M. I., Ashraf, S. & Intisar, A. Comparative study of different activation treatments for the preparation of activated carbon: A mini-review. Sci. Prog. 100, 299–312 (2017).

Article CAS PubMed Google Scholar

Mura, M. J., Behr, S., Bres, E. F. & Voegel, J. C. Dynamic processes at the liquid/solid interface in the albumin/apatite system. Clin. Mater. 5, 285–295. https://doi.org/10.1016/0267-6605(90)90028-T (1990).

Article CAS Google Scholar

Noubactep, C. & Sauter, M. Evaluation of the effects of shaking intensity on the process of methylene blue discoloration by metallic iron. J. Hazard. Mater. 169, 1005–1011 (2013).

Article Google Scholar

Tien, C. Introduction to Adsorption: Basics, Analysis, and Applications (Elsevier, 2019).

Google Scholar

Eduation, V. A. L. U. (vlab.amrita.edu, 2011).

Barbato, M. & Bruno, C. Molecular Physics and Hypersonic Flows 139–160 (Springer, 1996).

Book Google Scholar

Pourhakkak, P., Taghizadeh, A., Taghizadeh, M., Ghaedi, M. & Haghdoust, S. Fundamentals of Adsorption Technology in Adsorption: Fundamental Processes and Applications Vol. 33 (Academic Press, 2021).

Google Scholar

Rezania, S., Taib, S. M., Md Din, M. F., Dahalan, F. A. & Kamyab, H. Comprehensive review on phytotechnology: Heavy metals removal by diverse aquatic plants species from wastewater. J. Hazard. Mater. 318, 587–599. https://doi.org/10.1016/j.jhazmat.2016.07.053 (2016).

Article CAS PubMed Google Scholar

Vymazal, J. Concentration is not enough to evaluate accumulation of heavy metals and nutrients in plants. Sci. Total Environ. 544, 495–498. https://doi.org/10.1016/j.scitotenv.2015.12.011 (2016).

Article CAS PubMed Google Scholar

Kumari, M. & Tripathi, B. D. Efficiency of Phragmites australis and Typha latifolia for heavy metal removal from wastewater. Ecotoxicol. Environ. Saf. 112, 80–86. https://doi.org/10.1016/j.ecoenv.2014.10.034 (2015).

Article CAS PubMed Google Scholar

Rahbar, A. et al. Experimental data of biomaterial derived from Malva sylvestris and charcoal tablet powder for Hg2+ removal from aqueous solutions. Data Brief 8, 132–135. https://doi.org/10.1016/j.dib.2016.05.032 (2016).

Article PubMed PubMed Central Google Scholar

Emara, M. M., Farag, R. S., Mubarak, M. F. & Ali, S. K. Synthesis of core–shell activated carbon/CaO composite from Ficus Nitida leaves, as an efficient adsorbent for removal of methylene blue. Nanotechnol. Environ. Eng. 5, 8. https://doi.org/10.1007/s41204-020-00088-8 (2020).

Article CAS Google Scholar

Liu, Z., Khan, T. A., Islam, M. A. & Tabrez, U. A review on the treatment of dyes in printing and dyeing wastewater by plant biomass carbon. Bioresour. Technol. 354, 127168. https://doi.org/10.1016/j.biortech.2022.127168 (2022).

Article CAS PubMed Google Scholar

Ali, H., Khan, E. & Sajad, M. A. Phytoremediation of heavy metals—Concepts and applications. Chemosphere 91, 869–881. https://doi.org/10.1016/j.chemosphere.2013.01.075 (2013).

Article CAS PubMed Google Scholar

Ansari, A. A., Naeem, M., Gill, S. S. & AlZuaibr, F. M. Phytoremediation of contaminated waters: An eco-friendly technology based on aquatic macrophytes application. Egypt. J. Aquat. Res. 46, 371–376. https://doi.org/10.1016/j.ejar.2020.03.002 (2020).

Article Google Scholar

Galal, T. M. et al. Taif’s rose (Rosa damascena Mill var. trigentipetala) wastes are a potential candidate for heavy metals remediation from agricultural soil. Agriculture 12, 1319. https://doi.org/10.3390/agriculture12091319 (2022).

Article CAS Google Scholar

Török, A., Gulyás, Z., Szalai, G., Kocsy, G. & Majdik, C. Phytoremediation capacity of aquatic plants is associated with the degree of phytochelatin polymerization. J. Hazard. Mater. 299, 371–378. https://doi.org/10.1016/j.jhazmat.2015.06.042 (2015).

Article CAS PubMed Google Scholar

Siddiqui, S. H., Uddin, M. K., Isaac, R. & Aldosari, O. F. An effective biomass for the adsorption of methylene blue dye and treatment of river water. Adsorpt. Sci. Technol. 2022, 4143138. https://doi.org/10.1155/2022/4143138 (2022).

Article CAS Google Scholar

Al-Baldawi, I. A., Abdullah, S. R. S., Anuar, N. & Hasan, H. A. Phytotransformation of methylene blue from water using aquatic plant (Azolla pinnata). Environ. Technol. Innov. 11, 15–22. https://doi.org/10.1016/j.eti.2018.03.009 (2018).

Article Google Scholar

Boulos, L. Flora of Egypt (Alismataceae–Orchidaceae) (Al Hadara Publishing, 2005).

Google Scholar

Mahmoud, W. F. Population biology of Vossia cuspidata (Roxb.) Griff. in Egypt Ph.D. thesis, Cairo University (2020).

Galal, T. M., Gharib, F. A., Ghazi, S. M. & Mansour, K. H. Phytostabilization of heavy metals by the emergent macrophyte Vossia cuspidata (Roxb.) Griff.: A phytoremediation approach. Int. J. Phytorem. 19, 992–999. https://doi.org/10.1080/15226514.2017.1303816 (2017).

Article CAS Google Scholar

Galal, T. M., Eid, E. M., Dakhil, M. A. & Hassan, L. M. Bioaccumulation and rhizofiltration potential of Pistia stratiotes L. for mitigating water pollution in the Egyptian wetlands. Int. J. Phytorem. 20, 440–447. https://doi.org/10.1080/15226514.2017.1365343 (2018).

Article CAS Google Scholar

Farahat, E. A., Mahmoud, W. F. & Fahmy, G. M. Seasonal variations of heavy metals in water, sediment, and organs of Vossia cuspidata (Roxb.) Griff. in River Nile ecosystem: Implication for phytoremediation. Environ. Sci. Pollut. Res. 28, 32626–32633. https://doi.org/10.1007/s11356-021-13033-2 (2021).

Article CAS Google Scholar

Farahat, E. A. et al. Prediction models for evaluating the uptake of heavy metals by the invasive grass Vossia cuspidata (Roxb.) griff. in the river Nile, Egypt: A biomonitoring approach. Sustainability 13, 10558. https://doi.org/10.3390/su131910558 (2021).

Article CAS Google Scholar

Tang, C. et al. Comparison of the removal and adsorption mechanisms of cadmium and lead from aqueous solution by activated carbons prepared from Typha angustifolia and Salix matsudana. RSC Adv. 7, 16092–16103. https://doi.org/10.1039/c6ra28035h (2017).

Article CAS Google Scholar

Dilekoğlu, M. F. Malachite green adsorption from aqueous solutions onto biochar derived from sheep manure: Adsorption kinetics, isotherm, thermodynamic, and mechanism. Int. J. Phytorem. 24, 436–446. https://doi.org/10.1080/15226514.2021.1951656 (2022).

Article CAS Google Scholar

Azizian, S. Kinetic models of sorption: A theoretical analysis. J. Colloid Interface Sci. 276, 47–52. https://doi.org/10.1016/j.jcis.2004.03.048 (2004).

Article CAS PubMed Google Scholar

Weber, W. J. & Morris, J. C. Kinetics of adsorption on carbon from solution. J. Sanit. Eng. Div. 89, 31–59. https://doi.org/10.1061/JSEDAI.0000430 (1963).

Article Google Scholar

Lagergren, S. Zur theorie der sogenannten adsorption geloster stoffe. Kungliga svenska vetenskapsakademiens. Handlingar 24, 1–39 (1898).

Google Scholar

Revellame, E. D., Fortela, D. L., Sharp, W., Hernandez, R. & Zappi, M. E. Adsorption kinetic modeling using pseudo-first order and pseudo-second order rate laws: A review. Clean. Eng. Technol. 1, 100032. https://doi.org/10.1016/j.clet.2020.100032 (2020).

Article Google Scholar

Ho, Y. S. & McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 34, 451–465. https://doi.org/10.1016/S0032-9592(98)00112-5 (1999).

Article CAS Google Scholar

Lin, K., Pan, J., Chen, Y., Cheng, R. & Xu, X. Study the adsorption of phenol from aqueous solution on hydroxyapatite nanopowders. J. Hazard. Mater. 161, 231–240. https://doi.org/10.1016/j.jhazmat.2008.03.076 (2009).

Article CAS PubMed Google Scholar

Freundlich, H. M. F. Over the adsorption in solution. J. Phys. Chem. 57, 385–471 (1906).

CAS Google Scholar

Ayawei, N., Ebelegi, A. N. & Wankasi, D. Modelling and interpretation of adsorption isotherms. J. Chem. 2017, 3039817. https://doi.org/10.1155/2017/3039817 (2017).

Article CAS Google Scholar

Kajjumba, G. W., Emik, S., Öngen, A. & Aydın, H. K. Ö. S. Advanced Sorption Process Applications (IntechOpen, 2018).

Google Scholar

Hassan, M. A., Mohammad, A. M., Salaheldin, T. A. & El-Anadouli, B. E. A promising hydroxyapatite/graphene hybrid nanocomposite for methylene blue dye’s removal in wastewater treatment. Int. J. Electrochem. Sci. 13, 8222–8240. https://doi.org/10.20964/2018.08.77 (2018).

Article CAS Google Scholar

Mohammad, A. M., Salah Eldin, T. A., Hassan, M. A. & El-Anadouli, B. E. Efficient treatment of lead-containing wastewater by hydroxyapatite/chitosan nanostructures. Arab. J. Chem. 10, 683–690. https://doi.org/10.1016/j.arabjc.2014.12.016 (2017).

Article CAS Google Scholar

Salah, T. A., Mohammad, A. M., Hassan, M. A. & El-Anadouli, B. E. Development of nano-hydroxyapatite/chitosan composite for cadmium ions removal in wastewater treatment. J. Taiwan Inst. Chem. Eng. 45, 1571–1577. https://doi.org/10.1016/j.jtice.2013.10.008 (2014).

Article CAS Google Scholar

Bhattacharyya, K. G. & Gupta, S. S. Adsorptive accumulation of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) from water on montmorillonite: Influence of acid activation. J. Colloid Interface Sci. 310, 411–424. https://doi.org/10.1016/j.jcis.2007.01.080 (2007).

Article CAS PubMed Google Scholar

Tempkin, M. J. & Pyzhev, V. Kinetics of ammonia synthesis on promoted iron catalysts. Acta Physicochim. U.R.S.S. 12, 217–222 (1940).

Google Scholar

Hutson, N. D. & Yang, R. T. Theoretical basis for the Dubinin-Radushkevitch (D-R) adsorption isotherm equation. Adsorption 3, 189–195. https://doi.org/10.1007/bf01650130 (1997).

Article CAS Google Scholar

Dubinin, M. M. & Radushkevich, L. V. Equation of the characteristic curve of activated charcoal. Proc. Acad. Sci. Phys. Chem. U.S.S.R. 55, 331–333 (1947).

Google Scholar

Galal, T. M., Gharib, F. A., Al-Yasi, H. M., Mansour, K. H. & Hassan, M. M. Evaluation of the nutrient status and forage quality of the hippo grass (Vossia cuspidata (Roxb.) Griff.) along Ismailia canal, Egypt. J. Freshw. Ecol. 36, 63–76. https://doi.org/10.1080/02705060.2020.1853624 (2021).

Article CAS Google Scholar

Shooto, N. D., Naidoo, E. B. & Maubane, M. Sorption studies of toxic cations on ginger root adsorbent. J. Ind. Eng. Chem. 76, 133–140. https://doi.org/10.1016/j.jiec.2019.02.027 (2019).

Article CAS Google Scholar

Shooto, N. D., Thabede, P. M. & Naidoo, E. B. Simultaneous adsorptive study of toxic metal ions in quaternary system from aqueous solution using low cost black cumin seeds (Nigella sativa) adsorbents. S. Afr. J. Chem. Eng. 30, 15–27. https://doi.org/10.1016/j.sajce.2019.07.002 (2019).

Article Google Scholar

Alshehri, A. A. & Malik, M. A. Biogenic fabrication of ZnO nanoparticles using Trigonella foenum-graecum (Fenugreek) for proficient photocatalytic degradation of methylene blue under UV irradiation. J. Mater. Sci. Mater. Electron. 30, 16156–16173. https://doi.org/10.1007/s10854-019-01985-8 (2019).

Article CAS Google Scholar

Mashkoor, F., Nasar, A., Inamuddin, & Asiri, A. M. Exploring the reusability of synthetically contaminated wastewater containing crystal violet dye using Tectona grandis sawdust as a very low-cost adsorbent. Sci. Rep. 8, 3. https://doi.org/10.1038/s41598-018-26655-3 (2018).

Article CAS Google Scholar

Mashkoor, F. & Nasar, A. Polyaniline/Tectona grandis sawdust: A novel composite for efficient decontamination of synthetically polluted water containing crystal violet dye. Groundw. Sustain. Dev. 8, 390–401. https://doi.org/10.1016/j.gsd.2018.12.008 (2019).

Article Google Scholar

Bello, O. S., Adegoke, K. A., Sarumi, O. O. & Lameed, O. S. Functionalized locust bean pod (Parkia biglobosa) activated carbon for Rhodamine B dye removal. Heliyon 5, e02323. https://doi.org/10.1016/j.heliyon.2019.e02323 (2019).

Article PubMed PubMed Central Google Scholar

FTIR Functional Group Table with Search—InstaNANO. https://instanano.com/all/characterization/ftir/ftir-functional-group-search/ (2023).

Saikia, B. J., Parthasarathy, G., Borah, R. R. & Borthakur, R. Raman and FTIR spectroscopic evaluation of clay minerals and estimation of metal contaminations in natural deposition of surface sediments from Brahmaputra river. Int. J. Geosci. 7, 873–883 (2016).

Article CAS Google Scholar

Martsouka, F. et al. Evaluation of the antimicrobial protection of pharmaceutical kaolin and talc modified with copper and zinc. Materials 14, 1173 (2021).

Article CAS PubMed PubMed Central Google Scholar

Uddin, M. K. & Nasar, A. Walnut shell powder as a low-cost adsorbent for methylene blue dye: Isotherm, kinetics, thermodynamic, desorption and response surface methodology examinations. Sci. Rep. 10, 3. https://doi.org/10.1038/s41598-020-64745-3 (2020).

Article CAS Google Scholar

Amin, M. T., Alazba, A. A. & Shafiq, M. Successful application of eucalyptus camdulensis biochar in the batch adsorption of crystal violet and methylene blue dyes from aqueous solution. Sustainability 13, 3600. https://doi.org/10.3390/su13073600 (2021).

Article CAS Google Scholar

Iwuozor, K. O., Ighalo, J. O., Ogunfowora, L. A., Adeniyi, A. G. & Igwegbe, C. A. An empirical literature analysis of adsorbent performance for methylene blue uptake from aqueous media. J. Environ. Chem. Eng. 9, 105658. https://doi.org/10.1016/j.jece.2021.105658 (2021).

Article CAS Google Scholar

Shelke, B. N., Jopale, M. K. & Kategaonkar, A. H. Exploration of biomass waste as low cost adsorbents for removal of methylene blue dye: A review. J. Indian Chem. Soc. 99, 100530. https://doi.org/10.1016/j.jics.2022.100530 (2022).

Article CAS Google Scholar

Rafatullah, M., Sulaiman, O., Hashim, R. & Ahmad, A. Adsorption of methylene blue on low-cost adsorbents: A review. J. Hazard Mater. 177, 70–80. https://doi.org/10.1016/j.jhazmat.2009.12.047 (2010).

Article CAS PubMed Google Scholar

Rao, V. B. & Rao, S. R. M. Adsorption studies on treatment of textile dyeing industrial effluent by flyash. Chem. Eng. J. 116, 77–84 (2006).

Article CAS Google Scholar

Aygün, A., Yenisoy-Karakaş, S. & Duman, I. Production of granular activated carbon from fruit stones and nutshells and evaluation of their physical, chemical and adsorption properties. Microporous Mesoporous Mater. 66, 189–195 (2003).

Article Google Scholar

Kavitha, D. & Namasivayam, C. Experimental and kinetic studies on methylene blue adsorption by coir pith carbon. Bioresour. Technol. 98, 14–21 (2007).

Article CAS PubMed Google Scholar

Wu, F.-C. & Tseng, R.-L. High adsorption capacity NaOH-activated carbon for dye removal from aqueous solution. J. Hazard Mater. 152, 1256–1267 (2008).

Article CAS PubMed Google Scholar

Tseng, R.-L., Tseng, S.-K. & Wu, F.-C. Preparation of high surface area carbons from Corncob with KOH etching plus CO2 gasification for the adsorption of dyes and phenols from water. Colloids Surf. A Physicochem. Eng. Asp. 279, 69–78 (2006).

Article CAS Google Scholar

Ghosh, D. & Bhattacharyya, K. G. Adsorption of methylene blue on kaolinite. Appl. Clay Sci. 20, 295–300 (2002).

Article CAS Google Scholar

Gürses, A. et al. Determination of adsorptive properties of clay/water system: Methylene blue sorption. J. Colloid Interface Sci. 269, 310–314 (2004).

Article PubMed Google Scholar

Chakrabarti, S. & Dutta, B. K. On the adsorption and diffusion of methylene blue in glass fibers. J. Colloid Interface Sci. 286, 807–811 (2005).

Article CAS PubMed Google Scholar

Ncibi, M. et al. Biosorptive uptake of methylene blue using Mediterranean green alga Enteromorpha spp.. J. Hazard. Mater. 170, 1050–1055 (2009).

Article CAS PubMed Google Scholar

Cengiz, S. & Cavas, L. Removal of methylene blue by invasive marine seaweed: Caulerpa racemosa var. cylindracea. Bioresour. Technol. 99, 2357–2363 (2008).

Article CAS PubMed Google Scholar

Fu, Y. & Viraraghavan, T. Removal of a dye from an aqueous solution by the fungus Aspergillus niger. Water Qual. Res. J. 35, 95–112 (2000).

Article CAS Google Scholar

Batzias, F. & Sidiras, D. Dye adsorption by calcium chloride treated beech sawdust in batch and fixed-bed systems. J. Hazard. Mater. 114, 167–174 (2004).

Article CAS PubMed Google Scholar

Batzias, F., Sidiras, D., Schroeder, E. & Weber, C. Simulation of dye adsorption on hydrolyzed wheat straw in batch and fixed-bed systems. Chem. Eng. J. 148, 459–472 (2009).

Article CAS Google Scholar

Bhattacharyya, K. G. & Sharma, A. Kinetics and thermodynamics of methylene blue adsorption on neem (Azadirachta indica) leaf powder. Dyes Pigm. 65, 51–59 (2005).

Article CAS Google Scholar

Lee, C., Low, K. & Chow, S. Chrome sludge as an adsorbent for colour removal. Environ. Technol. 17, 1023–1028 (1996).

Article CAS Google Scholar

Kumar, K. Adsorption isotherm for basic dye onto low cost adsorbents. Res. J. Chem. Environ. 6, 61–65 (2002).

CAS Google Scholar

Krishnappa, B., Saravu, S., Shivanna, J. M., Naik, M. & Hegde, G. Fast and effective removal of textile dyes from the wastewater using reusable porous nano-carbons: a study on adsorptive parameters and isotherms. Environ. Sci. Pollut. Res. 29(52), 79067–79081 (2022).

Mahmoodi, N. M. & Mokhtari-Shourijeh, Z. Modified poly(vinyl alcohol)-triethylenetetramine nanofiber by glutaraldehyde: Preparation and dye removal ability from wastewater. Desalin. Water Treat. 57, 20076–20083. https://doi.org/10.1080/19443994.2015.1109562 (2016).

Article CAS Google Scholar

Pathania, D., Sharma, S. & Singh, P. Removal of methylene blue by adsorption onto activated carbon developed from Ficus carica bast. Arab. J. Chem. 10, S1445–S1451. https://doi.org/10.1016/j.arabjc.2013.04.021 (2017).

Article CAS Google Scholar

Samarbaf, S., Tahmasebi Birgani, Y., Yazdani, M. & Babaei, A. A. A comparative removal of two dyes from aqueous solution using modified oak waste residues: Process optimization using response surface methodology. J. Ind. Eng. Chem. 73, 67–77. https://doi.org/10.1016/j.jiec.2018.12.011 (2019).

Article CAS Google Scholar

Hussain, S., Anjali, K. P., Hassan, S. T. & Dwivedi, P. B. Waste tea as a novel adsorbent: A review. Appl. Water Sci. 8, 5. https://doi.org/10.1007/s13201-018-0824-5 (2018).

Article CAS Google Scholar

Prasanna Kumar, Y., King, P. & Prasad, V. S. R. K. Adsorption of zinc from aqueous solution using marine green algae—Ulva fasciata sp.. Chem. Eng. J. 129, 161–166. https://doi.org/10.1016/j.cej.2006.10.023 (2007).

Article CAS Google Scholar

Esmaeili, H. & Foroutan, R. Adsorptive behavior of methylene blue onto sawdust of sour lemon, date palm, and eucalyptus as agricultural wastes. J. Dispers. Sci. Technol. 40, 990–999. https://doi.org/10.1080/01932691.2018.1489828 (2019).

Article CAS Google Scholar

Soliman, N. K., Moustafa, A. F., Aboud, A. A. & Halim, K. S. A. Effective utilization of Moringa seeds waste as a new green environmental adsorbent for removal of industrial toxic dyes. J. Mater. Res. Technol. 8, 1798–1808. https://doi.org/10.1016/j.jmrt.2018.12.010 (2019).

Article CAS Google Scholar

Elmorsi, T. M. Equilibrium isotherms and kinetic studies of removal of methylene blue dye by adsorption onto miswak leaves as a natural adsorbent. J. Environ. Prot. 2, 817–827 (2011).

Article CAS Google Scholar

Shen, K. & Gondal, M. A. Removal of hazardous Rhodamine dye from water by adsorption onto exhausted coffee ground. J. Saudi Chem. Soc. 21, S120–S127. https://doi.org/10.1016/j.jscs.2013.11.005 (2017).

Article CAS Google Scholar

Piri, F., Mollahosseini, A., Khadir, A. & Milani Hosseini, M. Enhanced adsorption of dyes on microwave-assisted synthesized magnetic zeolite-hydroxyapatite nanocomposite. J. Environ. Chem. Eng. 7, 103338. https://doi.org/10.1016/j.jece.2019.103338 (2019).

Article CAS Google Scholar

Neina, D. The role of soil pH in plant nutrition and soil remediation. Appl. Environ. Soil Sci. 2019, 4869. https://doi.org/10.1155/2019/5794869 (2019).

Article CAS Google Scholar

Download references

The authors are grateful to Ms. Menna Rashad and Ms. Ahlam Zaki for their assistance in the early stages of running the adsorption experiments.

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Botany and Microbiology Department, Cairo University, Giza, 12613, Egypt

Hossam E. A. Awad

Chemistry Department, Faculty of Science, Cairo University, Cairo, 12613, Egypt

Ahmad M. Mohammad

Botany and Microbiology Department, Helwan University, Cairo, 11795, Egypt

Emad A. Farahat

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

All authors contributed critically to the drafts and gave final approval for publication.

Correspondence to Ahmad M. Mohammad.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Awad, H.E.A., Mohammad, A.M. & Farahat, E.A. Potential use of dry powder of Vossia cuspidata (Roxb.) Griff. rhizomes and leaves in methylene blue dye remediation. Sci Rep 13, 11073 (2023). https://doi.org/10.1038/s41598-023-37987-0

Download citation

Received: 26 January 2023

Accepted: 30 June 2023

Published: 08 July 2023

DOI: https://doi.org/10.1038/s41598-023-37987-0

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.