Next Article in Journal
Solidification/Stabilization of Arsenic-Containing Tailings by Steel Slag-Based Binders with High Efficiency and Low Carbon Footprint
Next Article in Special Issue
Preparation of Activated Carbon Derived from Jordanian Olive Cake and Functionalized with Cu/Cu2O/CuO for Adsorption of Phenolic Compounds from Olive Mill Wastewater
Previous Article in Journal
Bone Tissue Regeneration by Collagen Scaffolds with Different Calcium Phosphate Coatings: Amorphous Calcium Phosphate and Low-Crystalline Apatite
Previous Article in Special Issue
The Role of Carbon Nanotube Pretreatments in the Adsorption of Benzoic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bathurst Burr (Xanthium spinosum) Powder—A New Natural Effective Adsorbent for Crystal Violet Dye Removal from Synthetic Wastewaters

Faculty of Industrial Chemistry and Environmental Engineering, Politehnica University Timisoara, Bd. V. Parvan, No. 6, 300223 Timisoara, Romania
*
Authors to whom correspondence should be addressed.
Materials 2021, 14(19), 5861; https://doi.org/10.3390/ma14195861
Submission received: 23 August 2021 / Revised: 29 September 2021 / Accepted: 5 October 2021 / Published: 7 October 2021
(This article belongs to the Special Issue Materials for Waste Water Treatment)

Abstract

:
A new natural adsorbent material, Bathurst burr powder, was used to remove crystal violet dye from synthetic wastewaters. Particle size distribution and SEM and FTIR analyses were performed to characterize it. The effect of the operational adsorption process parameters (pH, ionic strength, initial dye concentration, adsorbent dose, contact time, temperature) onto the adsorption process was evaluated in a batch system. Equilibrium, kinetic, and thermodynamic studies were performed in order to understand the adsorption process. Taguchi method and ANOVA test were used to optimize the dye adsorption conditions and to establish the percentage contribution of each factor, respectively. The accuracy of the Taguchi prediction method was analyzed by correlating the predicted dye removal efficiency with the experimentally determined one. The particle size distribution analysis showed that 82.15% of the adsorbent particles have an average size below 0.5 mm. The adsorption process followed the Langmuir isotherm and pseudo-second order kinetic model. Maximum adsorption capacity value (164.10 mg·g−1) was higher compared to many similar adsorbents. The process was endothermic, spontaneous, and favorably involving a physisorption mechanism. The Taguchi method showed that the most influential controllable factor was pH (65% contribution in adsorption efficiency) and the data analysis indicates a very good accuracy of the experimental design (R2 = 0.994). The obtained results demonstrated that Bathurst burr powder can be used as a cheap and efficient adsorbent for crystal violet dye removal from aqueous solution.

1. Introduction

Crystal violet is a cationic dye used as a textile colorant, paper dye, biological stain, and as veterinary animal drug. It has high toxicity and can cause digestive tract irritation, kidney failure, and cancer [1,2,3,4,5]. Its presence in the aquatic environment, even in low concentrations, leads to plants photosynthesis process disorder. It has high chemical stability, high solubility, and low biodegradability. It is therefore necessary to remove this stain from residual effluents [1,2,3].
Many methods have been used to remove the cationic dyes (inclusive crystal violet) from wastewater and dye solutions. Coagulation, flocculation, electrocoagulation, ozonation and catalytic ozonation, photochemical and photocatalytic degradation, chemical and electrochemical oxidation, membrane processes (ultrafiltration, reverse osmosis), biodegradation, ion exchange, and adsorption have been mentioned in the literature as being effective for this purpose [1,3,4,5,6].
In recent years, the adsorption process has been used more and more often by researchers to achieve the removal of dyes from wastewaters. This is due to the fact that adsorption has many technical and economic advantages such as: high selectivity and efficiency, simplicity and flexibility in operation, good applicability, possibility to reuse the adsorbent, and relatively low costs [4,5,6].
The process becomes extremely cost-effective by using cheap materials that can be found in abundance in nature [7,8,9,10,11]. Several low-cost natural materials have been used for crystal violet dye removal from aqueous solutions: cedar cones [11], Calotropis procera leaf [12], Laminaria japonica [13], rice bran [13], wheat bran [13], coniferous pinus bark powder [14], sugarcane fiber [15], sawdust [15], coir pith [15], peel of Cucumis sativa fruit [16], jackfruit leaf powder [17], pineapple leaf powder [18], papaya seeds [19], breadfruit skin [8], water hyacinth root powder [20], Eragrostis plana nees [7], pará chestnut husk [21], date palm leaves powder [22], corn stalk [23], Ocotea puberula bark powder [5], Moringa oleifera pod husk [24], and almond shells [25].
Bathurst burr (Xanthium spinosum) is an herbaceous plant, up to 1 meter tall, with a richly branched stem and full of thorns. It grows especially in hilly and plain areas and also in low mountainous regions. Originally from South America, it has spread widely in Europe, Australia, parts of Africa, Asia, and North America, and is a very invasive plant. It grows on pastures, roadsides, fences, and abandoned lands and has an extraordinary resistance to drought, pollution, and any aggressive environmental factor. It is used in traditional medicine based on its: decongestant, diuretic, anti-inflammatory, disinfectant, diuretic, antidiabetic, and antitumor properties [26,27]. The seeds of the plant contain carboxyatractyloside, which is poisonous to livestock, but animals usually avoid this plant because of its thorns, which can cause them unpleasant wounds [28,29].
The aim of this research study was to remove crystal violet dye from synthetic wastewaters using the Bathurst burr powder as an adsorbent material. The effect of pH, contact time, adsorbent dose, initial dye concentration, temperature, and ionic strength onto the adsorption process was evaluated in the batch system. In order to understand the adsorption mechanism, different isotherms, kinetics, and thermodynamics studies were performed. The Taguchi method was used for adsorption conditions optimization. In addition, the ANOVA analysis was used to assess the result and to determine the contribution percentage of each parameter on the dye removal process.

2. Materials and Methods

In order to obtain the adsorbent material, the aerial part of Bathurst burr mature plants were purchased from the company StefMar (Ramnicu Valcea, Romania) whose activity is the processing and packaging of medicinal and aromatic plants. The plants were washed with distilled water and then first dried at room temperature for 3 days and then in an air oven at 105 °C for 24 h. After drying, they were electrically grounded. The obtained powder was passed through a 2 mm sieve.
SEM analysis and FTIR spectroscopy were performed to characterize the adsorbent material using a Quanta FEG 250 (FEI, Eindhoven, The Netherlands) scanning electron microscope (1600× magnitude) and a Shimadzu Prestige-21 FTIR (Shimadzu, Kyoto, Japan) spectrophotometer. Analyses were compared before and after the dye adsorption process.
In batch adsorption experiments, the influence of various factors affecting the process was studied, at a constant stirring intensity provided by a shaker. The pH was adjusted using 0.1 N NaOH and HCl solutions. The influence of ionic strength was investigated using NaCl as background electrolyte. The crystal violet concentration was measured with a Specord 200 PLUS UV-VIS (Analytik Jena, Jena, Germany) spectrophotometer at a wavelength of 590 nm. Each test was conducted using three independent replicates.
The crystal violet amounts adsorbed at equilibrium (qe) and the dye removal percentage R(%) were calculated with equations (1) and (2) [1,4,7]:
q e = ( C 0 C e ) V m ,
R ( % ) = ( C 0 C e ) C 0 100 ,
where: C0 is the initial crystal violet concentration (mg·L−1), Ce is the crystal violet equilibrium concentration (mg·L−1), V is the solution volume (L), and m is the mass of adsorbent (g).
The optimum conditions for crystal violet adsorption were established using Taguchi (L27) experimental design. To achieve this goal, the effect of six factors (at three levels) on dye removal efficiency was followed. These controllable factors and their levels are detailed in Table 1. For the experimental design, the Taguchi method uses an orthogonal matrix (OA) evaluating the signal/noise ratio (S/N) using the “larger is the better” option [30,31,32]. In order to evaluate the results of the Taguchi method and to establish the percentage contribution of each factor to the efficiency of crystal violet removal, an analysis of variance (ANOVA) was realized [30,32]. The necessary calculations were performed with the Minitab 19 Software (version 19.1.1, Minitab LLC, State College, PA, USA).
The desorption study was realized in batch mode, at constant stirring for 2 h, using different desorbing agents (distilled water, 0.1 M HCl, and 0.1 M NaOH).

3. Results and Discussion

3.1. Adsorbent Characterization

The particle size distribution analysis of the Bathurst burr powder (Supplementary Material, Figure S1) shows that 82.15% of the particles have an average size bellow 0.5 mm and 50% of them are smaller than 0.348 mm.
Figure 1 shows the morphology of the adsorbent before and after adsorption. Before adsorption, the surface had many pores that provided a large adsorption surface and a large number of sites available for retaining dye molecules (Figure 1a). After adsorption, the surface of the adsorbent was saturated and covered with crystal violet molecules, which filled the pores and cavities during the process (Figure 1b).
The FTIR spectra of adsorbent before and after adsorption (Figure 2) indicate different bands specific for main functional group existing in cellulose (3448 cm−1, 2340 cm−1, 1630 cm−1, 1368 cm−1, and 1000 cm−1) [33,34,35,36,37] and hemicellulose (3448 cm−1, 1368 cm−1) [36,38,39,40]. These assignments are shown in Table 2. After adsorption, the appearance or disappearance of any band was not found. It only presents small changes of the wavenumber, suggesting that the physical interaction or ion exchange can intervene in the adsorption mechanism [41].

3.2. Effect of pH and Ionic Strength on Crystal Violet Adsorption

pH is an important parameter that affects the adsorbent surface charge [42]. The influence of this parameter on the adsorption capacity was followed in the pH range 2–12 (Figure 3a). The adsorption capacity increases with pH, reaching the highest values at basic conditions. A similar phenomenon was observed in other similar studies regarding the adsorption of crystal violet on low-cost adsorbents [8,9,43,44]. The adsorbent materials have different functional groups on the surface. At lower pH these groups will be protonated and the charge on the adsorbent surface will be predominantly positive, therefore, the adsorption process will be more difficult due to electrostatic repulsion with cationic dye molecules. For higher pH values, these functional groups will be deprotonated and the charge on the adsorbent surface will be predominantly negative. This phenomenon favors the adsorption process because there is a strong electrostatic attraction between negatively charged adsorbent surface and the dye cations [18,22,45]. The adsorption capacity on the pH range 8–12 remains practically constant, suggesting that other mechanisms besides electrostatic attraction may be involved in the adsorption process [46,47].
Another important factor that can influence the adsorption process is the ionic strength. The effect of this factor on the adsorption capacity is presented in Figure 3b. The decrease of adsorption capacity as the ionic strength increases is based on the competitive effect between the dye and sodium ions in occupying the available adsorption sites on the adsorbent surface [45,46]. A similar influence of ionic strength on the crystal violet adsorption process has been reported in other previous scientific articles [43,48,49].

3.3. Effect of Adsorbent Dose and Initial Dye Concentration on Crystal Violet Adsorption

Figure 4a illustrates the adsorbent dose effect on the crystal violet adsorption process. The increase of adsorbent amount has a positive effect on the dye removal efficiency and a negative effect on the adsorption capacity. A similar phenomenon has been reported for this dye adsorbed using: eggshells [4], Ocotea puberula bark powder [5], pinus bark powder [14], lemongrass leaf combined with cellulose acetate [42], functionalized multi-walled carbon nanotubes [50], and inactive biomass of Diaporthe schini [51]. The increase of the adsorbent material dose leads to an increase of adsorption surface; therefore, removal efficiency of the dye will be higher [11,14,18,50]. The reported decrease of the adsorption capacity (qe) can be explained based on the assumption that even if the adsorption sites number increases, many of them remain unsaturated. Other phenomenon such as agglomeration of adsorbent material particles may also occur [4,23,50,51].
The adsorption capacity increases and the removal efficiency decreases as initial dye concentration increases, as shown in Figure 4b. Similar effects were mentioned for other adsorbents used to retain crystal violet: eggshells [4], Ananas comosus (pineapple) leaf powder [18], and corn stalk [23]. The increase of the adsorption capacity can be attributed to the fact that a high initial dye concentration leads to an increase of the concentration gradient between the dye solution and the adsorbent surface, leading to an increase of the driving force that favors the external mass transfer [4,11,18,51,52]. At the same time, the number of collisions between dye molecules and adsorbent particles material are favored, intensifying adsorption [53]. The unfavorable effect of increasing initial dye concentration upon removal efficiency derives from the fact that the active adsorption sites become saturated with the dye molecules accumulated during the process [4,18,50,54].

3.4. Equilibrum Study

Adsorption isotherms are used to describe the interactions between adsorbate and adsorbent that take place in the process. The calculated parameters using these isotherms provide information about the properties and affinities of the adsorbent material surface as well as about the adsorption mechanism [1,11,52].
Based on the data presented in Figure 4b (experimental results and working conditions), Langmuir and Freundlich isotherms, in the form of non-linear equations [7,8,43,55,56,57], were used for this purpose:
Langmuir isotherm non - linear equation : q e = q m K L C e 1 + K L C e ,
Freundlich isotherm non - linear equation : q e = K F C e 1 / n F ,
where: qm is the maximum absorption capacity (mg·g−1), KL and KF are the Langmuir and Freundlich constants, respectively, and 1/nF is an empirical constant indicating the intensity of adsorption.
Both isotherms curves for crystal violet adsorption are presented in the Supplementary Material, Figure S2.
In order to establish the best equations that describe the adsorption process, values of determination coefficient (R2), sum of square error (SSE), chi-square (χ2) and average relative error (ARE) were considered [57]. The criterion for their applicability was the higher value for R2 and the lower values for SSE, χ2, and ARE. These parameters were determined with equations (5)−(8) [57].
R 2 = 1 i = 1 n y i , e x p y i , m o d 2 i = 1 n y i , e x p y i , e x p ¯ 2 ,
S S E = i = 1 n y i , e x p y i , m o d 2 ,
χ 2 = i = 1 n y i , e x p y i , m o d 2 y i , m o d ,
A R E = 100 n i = 1 n y i , e x p y i , m o d y i , m o d ,
where: yi,exp is the experimental value, yi,mod is the modeled value, y i , e x p ¯ is the mean values, and n is the total amount of information.
The values of (R2), (SSE), (χ2) and (ARE) are summarized in Table 3. It can be concluded that Langmuir isotherm best describes the adsorption process.
The RL separation coefficient can be used to predict whether an adsorption system is favorable or unfavorable. It can be calculated using the following equation:
R L = 1 1 + K L C 0 ,
where: KL is the Langmuir constant, and C0 is the initial dye concentration (mg·L−1) [10,11,22].
The calculated value for RL (0.29) indicates a favorable adsorption.
The value of the maximum adsorption capacity obtained for Bathurst burr powder, 164.10 (mg·g−1), is higher than other similar adsorbents previously reported in the literature. Table 4 shows the values for the maximum adsorption capacities of several similar adsorbents used for the adsorption of crystal violet dye.

3.5. Kinetic and Thermodynamic Studies

The adsorption kinetics study provides information on the operating conditions and the kinetic parameters. These are the basis for predicting the adsorption rate and designing the adsorption processes [52,59].
The effect of contact time on crystal violet adsorption is shown in Figure 5a. The adsorption capacity increases with the contact time between the dye solution and the adsorbent material. After 30 min, the equilibrium is reached and the value of this parameter remains practically constant. In the first 15 min the increase is more pronounced because at the beginning of the process many free adsorption sites on the surface of the adsorbent are available to retain the dye, [8,9,24,42,49]. After 15 min the adsorption capacity increases more slowly, because more and more free sites on the adsorbent surface are occupied by dye molecules and repulsive forces between adsorbed and liquid phase molecules may occur [8,9,45,50]. The increase of the adsorption capacity stops when the equilibrium is reached, the whole adsorbent surface becomes saturated with dye, and the adsorption sites are all occupied [8,9,47].
The equilibrium times, reported in scientific literature, for different crystal violet adsorbents, were: 15 min for pecan pericarp [3]; 30 min for Artocarpus odoratissimus leaf-based cellulose [43] and graphene oxide intercalated montmorillonite nanocomposite [60]; 45 min for functionalized multi-walled carbon nanotubes [50]; 60 min for Moringa oleifera pod husk [24]; 90 min for NaOH-modified rice husk [45]; 120 min for Ocotea puberula bark powder [5], Terminalia arjuna sawdust [9], and pinus bark powder [14]; 150 min for yeast-treated peat [44]; 180 min for Eragrostis plana nees [7]; 240 min for lemongrass leaf combined with cellulose acetate [42]; and 420 min for crosslinked grafted xanthan gum [48].
The pseudo-first-order and pseudo-second-order models, in the form of non-linear equations [7,8,43,55,56,57] were employed to study the adsorption kinetics:
Pseudo - first - order model equation : q t = q e ( 1 e x p k 1 t ) ,
Pseudo - second - order model equation : q t = k 2 t q e 2 1 + k 2 t q e ,
where: qt is the crystal violet amount adsorbed at time t (mg·g−1), and k1 and k2 are the rate constants of pseudo-first-order kinetic model and the pseudo-second-order kinetic model, respectively.
Both kinetic model curves for crystal violet adsorption are presented in the Supplementary Material, Figure S3.
The pseudo-second-order kinetic model had the highest value for the determination coefficient (R2) and the lowest values for sum of square error (SSE), chi-square (χ2), and average relative error (ARE), (Table 5). Therefore, this model is the most suitable to characterize the process. A similar observation has been reported in other previous scientific articles [7,11,14,16,18,19,25].
Figure 5b illustrates the effect of temperature on the adsorption capacity, indicating that the adsorption process is endothermic in nature [61,62]. A similar aspect has been reported in other previous studies regarding crystal violet absorption [44,54]. The increase of the adsorption capacity can be explained based on the increasing mobility of the dye molecules due to the viscosity decrease with temperature [44,47,54,62].
Standard Gibbs free energy change, standard enthalpy change, and standard entropy change were calculated from experimental data obtained at temperatures of 285, 294, and 311 K using the equations described in literature [7,56]:
Δ G 0 = R T l n K L ,
l n K L = Δ S 0 R Δ H 0 R T ,
where: R is the universal gas constant, KL is the Langmuir constant, and T is the absolute temperature.
Thermodynamic parameters (Table 6) were determined from the slope and the intercept of ln KL versus 1/T plot (see Supplementary Material, Figure S4). The data analysis show that the adsorption process is endothermic, spontaneous, and favorable (ΔH0 > 0, ΔG0 < 0). The affinity of adsorbent for crystal violet and the increased randomness at the solid–solute interface (the degrees of freedom of the adsorbed species) are indicated by the positive value of ΔS0 [1,25,56]. Similar results were obtained by other researchers on low-cost adsorbent materials [7,14,25]. The physisorption is involved in the process (ΔH0 < 40 kJ mol−1) [25,44] and van der Waals interaction plays an important role in the physical adsorption (ΔH0 < 20 kJ mol−1) [63]. When standard Gibbs free energy change (ΔG0) ranges from −20 to −80 (kJ mol−1) the process is physisorption, intensified by the chemical effect [49].

3.6. Optimization Parameters of Adsorption Process Using Taguchi Approach

The Taguchi method was used to determine the optimal experimental conditions for obtaining the highest efficiency of dye removal from water, using a minimized number of experiments. According to this method, the experimental results are converted into a signal-to-noise (S/N) ratio, which is an indicator that describes at the same time the level of dispersion and degree of optimization, in relation to the desired value. Depending on the number of controllable parameters and their levels, the Taguchi method uses the proper and most representative orthogonal array that distributes the variables in a balanced way [30,31,32]. For six controllable factors at three levels, the classical full factorial design must use more experiments (36 = 729) to establish optimal conditions. In this case, the proper Taguchi orthogonal array is L27, which reduce the number of experiments to 27.
The obtained results for dye removal efficiency and the S/N ratios for each run are presented in Supplementary Material, Table S1. The S/N ratio rank was used to establish the order of the controllable factors’ significance (Table 7). The factor that had the least influence on the process was the temperature, while the factor with the greatest influence was the pH. The optimal conditions to obtain the highest efficiency of crystal violet removal are highlighted in Table 7. ANOVA results for signal-to-noise S/N ratios certify the same order of influence for the controllable factor. The percentage contribution of each factor is shown in Table 7. The predicted crystal violet removal efficiency was correlated with the experimentally determined one, and it was found that the accuracy of the prediction of the Taguchi method is very good, R2 being close to unity (Figure 6).

3.7. Desorption Study

The regeneration possibility of the used adsorbent was performed using three desorbing agents (0.1 M HCl, distilled water, and 0.1 M NaOH). HCl was found as the best desorption agent (see Supplementary Information, Figure S5) but the value of desorption efficiency was lower than 40%. Based on these results and taking into account that Bathurst burr powder is a cheap and easily available material and that desorption involves costs associated with the desorption reagents, it can be concluded that regeneration is not recommended.
The saturated adsorbent, based on its combustion properties as a vegetable material, can be used for direct combustion in specialized incinerators. Another new alternative is to use it as foaming agent for porous glass and ceramic materials based on the gases that result during the thermal synthesis process.

4. Conclusions

The natural adsorbent material, Bathurst burr powder, can be successfully used for crystal violet dye removal from aqueous solution. The removal efficiency and the adsorption capacity of this material are influenced by pH, contact time, initial dye concentration, adsorbent dose, and ionic strength. The adsorption process was found to follow Langmuir isotherm and pseudo-second order kinetic model. Maximum adsorption capacity value was higher compared to many similar adsorbents reported in the literature. The adsorption process is endothermic, spontaneous, and favorable. A physisorption mechanism, intensified by the chemical effect, is involved, and the van der Waals interaction plays an important role in the physical adsorption. The Taguchi method shows that the most influential controllable factor was the pH, while the least influential controllable factor was the initial dye concentration. The correlation between the predicted and the experimentally determined dye removal efficiency indicates a very good accuracy of the Taguchi experimental design.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma14195861/s1, Table S1: Experimental layout of L27 orthogonal array and results obtained for removal efficiency and S/N ratios, Figure S1: The particle size distribution of the Bathurst burr powder, Figure S2: Langmuir and Freundlich adsorption isotherms (non-linear forms) for the crystal violet adsorption onto Bathurst burr powder, Figure S3: Pseudo-first-order and pseudo-second-order kinetic models (non-linear forms) for the crystal violet adsorption onto Bathurst burr powder, Figure S4: Plot of ln KL vs. 1/T for the crystal violet adsorption onto Bathurst burr powder, Figure S5: The desorption efficiencies of crystal violet dye in different media.

Author Contributions

Conceptualization, G.M., C.V. and S.P.; methodology, G.M.; software, G.M. and C.V.; validation, G.M.; formal analysis, G.M., C.V. and S.P.; investigation, G.M., S.P. and S.B.; resources, G.M.; data curation, G.M.; writing—original draft preparation, G.M., C.V., S.P. and S.B.; writing—review and editing, G.M., C.V. and S.P.; visualization, G.M.; supervision, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the experimental data obtained are presented, in the form of table and/or figure, in the article and in the Supplementary materials.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aysu, T.; Küçük, M.M. Removal of crystal violet and methylene blue from aqueous solutions by activated carbon prepared from Ferula orientalis. Int. J. Environ. Sci. Technol. 2014, 12, 2273–2284. [Google Scholar] [CrossRef] [Green Version]
  2. Chahinez, H.-O.; Abdelkader, O.; Leila, Y.; Tran, H.N. One-stage preparation of palm petiole-derived biochar: Characterization and application for adsorption of crystal violet dye in water. Environ. Technol. Innov. 2020, 19, 100872. [Google Scholar] [CrossRef]
  3. Franco, D.S.P.; Fagundes, J.L.S.; Georgin, J.; Salaua, N.P.G.; Dotto, G.L. A mass transfer study considering intraparticle diffu-sion and axial dispersion for fixed-bed adsorption of crystal violet on pecan pericarp (Carya illinoensis). Chem. Eng. J. 2020, 397, 125423. [Google Scholar] [CrossRef]
  4. Chowdhury, S.; Chakraborty, S.; Das, P. Removal of Crystal Violet from Aqueous Solution by Adsorption onto Eggshells: Equilibrium, Kinetics, Thermodynamics and Artificial Neural Network Modeling. Waste Biomass- Valorization 2012, 4, 655–664. [Google Scholar] [CrossRef]
  5. Georgin, J.; Franco, D.S.P.; Netto, M.S.; Allasia, D.; Oliveira, M.L.S.; Dotto, G.L. Evaluation of Ocotea puberula bark powder (OPBP) as an effective adsorbent to uptake crystal violet from colored effluents: Alternative kinetic approaches. Environ. Sci. Pollut. Res. 2020, 27, 25727–25739. [Google Scholar] [CrossRef]
  6. Rosly, N.Z.; Abdullah, A.H.; Kamarudin, M.A.; Ashari, S.E.; Ahmad, S.A.A. Adsorption of Methylene Blue Dye by Calix[6]Arene-Modified Lead Sulphide (Pbs): Optimisation Using Response Surface Methodology. Int. J. Environ. Res. Public Health 2021, 18, 397. [Google Scholar] [CrossRef]
  7. Filho, A.C.D.; Mazzocato, A.C.; Dotto, G.L.; Thue, P.S.; Pavan, F.A. Eragrostis plana Nees as a novel eco-friendly adsorbent for removal of crystal violet from aqueous solutions. Environ. Sci. Pollut. Res. 2017, 24, 19909–19919. [Google Scholar] [CrossRef]
  8. Lim, L.B.L.; Priyantha, N.; Mansor, N.H.M. Artocarpus altilis (breadfruit) skin as a potential low-cost biosorbent for the removal of crystal violet dye: Equilibrium, thermodynamics and kinetics studies. Environ. Earth Sci. 2014, 73, 3239–3247. [Google Scholar] [CrossRef]
  9. Shakoor, S.; Nasar, A. Adsorptive decontamination of synthetic wastewater containing crystal violet dye by employing Ter-minalia arjuna sawdust waste. Groundw. Sustain Dev. 2018, 7, 30–38. [Google Scholar] [CrossRef]
  10. Shoukat, S.; Bhatti, H.N.; Iqbal, M.; Noreen, S. Mango stone biocomposite preparation and application for crystal violet ad-sorption: A mechanistic study. Micropor Mesopor Mater. 2017, 239, 180–189. [Google Scholar] [CrossRef]
  11. Zamouche, M.; Habib, A.; Saaidia, K.; Lehocine, M.B. Batch mode for adsorption of crystal violet by cedar cone forest waste. SN Appl. Sci. 2020, 2, 198. [Google Scholar] [CrossRef] [Green Version]
  12. Ali, H.; Muhammad, S.K. Biosorption of Crystal Violet from Water on Leaf Biomass of Calotropis procera. J. Environ. Sci. Technol. 2008, 1, 143–150. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, X.S.; Liu, X.; Wen, L.; Zhou, Y.; Jiang, Y.; Li, Z. Comparison of Basic Dye Crystal Violet Removal from Aqueous Solution by Low-Cost Biosorbents. Sep. Sci. Technol. 2008, 43, 3712–3731. [Google Scholar] [CrossRef]
  14. Ahmad, R. Studies on adsorption of crystal violet dye from aqueous solution onto coniferous pinus bark powder (CPBP). J. Hazard. Mater. 2009, 171, 767–773. [Google Scholar] [CrossRef] [PubMed]
  15. Parab, H.; Sudersanan, M.; Shenoy, N.; Pathare, T.; Vaze, B. Use of Agro-Industrial Wastes for Removal of Basic Dyes from Aqueous Solutions. CLEAN - Soil, Air, Water 2009, 37, 963–969. [Google Scholar] [CrossRef]
  16. Smitha, T.; Thirumalisamy, S.; Manonmani, S. Equilibrium and Kinetics Study of Adsorption of Crystal Violet onto the Peel of Cucumis sativa Fruit from Aqueous Solution. E-Journal Chem. 2012, 9, 1091–1101. [Google Scholar] [CrossRef] [Green Version]
  17. Das, P.; Chakraborty, S.; Chowdhury, S. Batch and continuous (fixed-bed column) biosorption of crystal violet by Artocarpus heterophyllus (jackfruit) leaf powder. Colloids Surf. B Colloid Surface B 2012, 92, 262–270. [Google Scholar] [CrossRef]
  18. Chakraborty, S.; Chowdhury, S.; Das, P. Insight into biosorption equilibrium, kinetics and thermodynamics of crystal violet onto Ananas comosus (pineapple) leaf powder. Appl. Water Sci. 2012, 2, 135–141. [Google Scholar] [CrossRef] [Green Version]
  19. Pavan, F.A.; Camacho, E.S.; Lima, E.C.; Dotto, G.L.; Branco, V.T.; Dias, S. Formosa papaya seed powder (FPSP): Preparation, characterization and application as an alternative adsorbent for the removal of crystal violet from aqueous phase. J. Environ. Chem. Eng. 2014, 2, 230–238. [Google Scholar] [CrossRef]
  20. Kulkarni, M.R.; Revanth, T.; Acharya, A.; Bhat, P. Removal of Crystal Violet dye from aqueous solution using water hyacinth: Equilibrium, kinetics and thermodynamics study. Resour. Technol. 2017, 3, 71–77. [Google Scholar] [CrossRef]
  21. Georgin, J.; Marques, B.S.; Peres, E.C.; Allasia, D.; Dotto, G.L. Biosorption of cationic dyes by Pará chestnut husk (Bertholletia excelsa). Water Sci. Technol. 2018, 77, 1612–1621. [Google Scholar] [CrossRef] [Green Version]
  22. Ghazali, A.; Shirani, M.; Semnania, A.; Zare-Shahabadic, V.; Nekoeiniad, M. Optimization of crystal violet adsorption onto Date palm leaves as a potent biosorbent from aqueous solutions using response surface methodology and ant colony. J. Environ. Chem. Eng. 2018, 6, 3942–3950. [Google Scholar] [CrossRef]
  23. Muhammad, U.L.; Zango, Z.U.; Kadir, H.A.; Usman, A. Crystal violet removal from aqueous solution using corn stalk bio-sorbent. Sci. World J. 2019, 14, 133–138. [Google Scholar]
  24. Keereerak, A.; Chinpa, W. A potential biosorbent from Moringa oleifera pod husk for crystal violet adsorption: Kinetics, iso-therms, thermodynamic and desorption studies. Sci. Asia. 2020, 46, 186–194. [Google Scholar] [CrossRef]
  25. Loulidi, I.; Boukhlifi, F.; Ouchabi, M.; Amar, A.; Jabri, M.; Kali, A.; Chraibi, S.; Hadey, C.; Aziz, F. Adsorption of Crystal Violet onto an Agricultural Waste Residue: Kinetics, Isotherm, Thermodynamics, and Mechanism of Adsorption. Sci. World, J. 2020, 2020, 1–9. [Google Scholar] [CrossRef] [PubMed]
  26. Holm, L.R.G.; Plucknett, D.L.; Pancho, J.V.; Herberger, J.P. The World’s Worst Weeds. Distribution and Biology; Krieger Pub. Co.: Malabar, FL, USA, 1977; p. 610. [Google Scholar]
  27. Raman, G.; Park, K.T.; Kim, J.-H.; Park, S. Characteristics of the completed chloroplast genome sequence of Xanthium spinosum: Comparative analyses, identification of mutational hotspots and phylogenetic implications. BMC Genom. 2020, 21, 1–14. [Google Scholar] [CrossRef]
  28. Global Invasive Species Database (GISD) 2021. Species Profile Xanthium Spinosum. Available online: http://www.iucngisd.org/gisd/species.php?sc=1347 (accessed on 23 September 2021).
  29. Amin, S.; Barkatullah; Khan, H. Pharmacology of Xanthium species. A review. J. Phytopharm. 2016, 5, 126–127. [Google Scholar] [CrossRef]
  30. Fernández-López, J.A.; Angosto, J.M.; Roca, M.J.; Miñarro, M.D. Taguchi design-based enhancement of heavy metals bioremoval by agroindustrial waste biomass from artichoke. Sci. Total. Environ. 2018, 653, 55–63. [Google Scholar] [CrossRef]
  31. Santra, D.; Joarder, R.; Sarkar, M. Taguchi design and equilibrium modeling for fluoride adsorption on cerium loaded cellulose nanocomposite bead. Carbohydr. Polym. 2014, 111, 813–821. [Google Scholar] [CrossRef]
  32. Zolgharnein, J.; Rastgordani, M. Optimization of simultaneous removal of binary mixture of indigo carmine and methyl orange dyes by cobalt hydroxide nano-particles through Taguchi method. J. Mol. Liq. 2018, 262, 405–414. [Google Scholar] [CrossRef]
  33. Senthamaraikannan, P.; Sanjay, M.R.; Bhat, K.S.; Padmaraj, N.H.; Jawaid, M. Characterization of natural cellulosic fiber frombark of Albizia amara. J. Nat. Fibers 2018, 1–8. [Google Scholar] [CrossRef]
  34. Tsuboi, M. Infrared spectrum and crystal structure of cellulose. J. Polym. Sci. 1957, 25, 159–171. [Google Scholar] [CrossRef]
  35. Karimi, S.; Tahir, P.M.; Karimi, A.; Dufresne, A.; Abdulkhani, A. Kenaf bast cellulosic fibers hierarchy: A comprehensive approach from micro to nano. Carbohydr. Polym. 2013, 101, 878–885. [Google Scholar] [CrossRef] [PubMed]
  36. Labbe, N.; Rials, T.G.; Kelley, S.S.; Cheng, Z.M.; Kim, J.Y.; Li, Y. FT-IR imaging and pyrolysis-molecular beam mass spec-trometry: New tools to investigate wood tissues. Wood Sci. Technol. 2005, 39, 61–76. [Google Scholar] [CrossRef]
  37. Liang, C.Y.; Marchessault, R.H. Infrared spectra of crystalline polysaccharides. II.Native celluloses in the region from 640 to 1700 cm−1. J. Polym. Sci. 1959, 39, 269–278. [Google Scholar] [CrossRef]
  38. Pardo, L.M.F.; Córdoba, A.G.; Galán, J.E.L. Characterization of hemicelluloses from leaves and tops of the CC 8475, CC 8592, and V 7151 varieties of sugarcane (Saccharum officinarum L.). DYNA 2019, 86, 98–107. [Google Scholar] [CrossRef] [Green Version]
  39. Kubovský, I.; Kačíková, D.; Kačík, F. Structural Changes of Oak Wood Main Components Caused by Thermal Modification. Polymers 2020, 12, 485. [Google Scholar] [CrossRef] [Green Version]
  40. Boukir, A.; Fellak, S.; Doumenq, P. Structural characterization of Argania spinosa Moroccan wooden artifacts during natural degradation progress using infrared spectroscopy (ATR-FTIR) and X-Ray diffraction (XRD). Heliyon 2019, 5, e02477. [Google Scholar] [CrossRef] [Green Version]
  41. Blázquez, G.; Martín-Lara, M.; Tenorio, G.; Calero, M. Batch biosorption of lead(II) from aqueous solutions by olive tree pruning waste: Equilibrium, kinetics and thermodynamic study. Chem. Eng. J. 2011, 168, 170–177. [Google Scholar] [CrossRef]
  42. Putri, K.N.A.; Keereerak, A.; Chinpa, W. Novel cellulose-based biosorbent from lemongrass leaf combined with cellulose acetate for adsorption of crystal violet. Int. J. Biol. Macromol. 2020, 156, 762–772. [Google Scholar] [CrossRef]
  43. Zaidi, N.A.H.M.; Lim, L.B.L.; Usman, A. Artocarpus odoratissimus leaf-based cellulose as adsorbent for removal of methyl violet and crystal violet dyes from aqueous solution. Cellulose 2018, 25, 3037–3049. [Google Scholar] [CrossRef]
  44. Zehra, T.; Priyantha, N.; Lim, L.B.L. Removal of crystal violet dye from aqueous solution using yeast-treated peat as adsorbent: Thermodynamics, kinetics, and equilibrium studies. Environ. Earth Sci. 2016, 75, 1–15. [Google Scholar] [CrossRef]
  45. Chakraborty, S.; Chowdhury, S.; Saha, P.D. Adsorption of crystal violet from aqueous solution onto NaOH-modified rice husk. Carbohydr. Polym. 2011, 86, 1533–1541. [Google Scholar] [CrossRef]
  46. Han, X.; Wang, W.; Ma, X. Adsorption characteristics of methylene blue onto low cost biomass material lotus leaf. Chem. Eng. J. 2011, 171, 1–8. [Google Scholar] [CrossRef]
  47. Pang, J.; Fu, F.; Ding, Z.; Lu, J.; Li, N.; Tang, B. Adsorption behaviors of methylene blue from aqueous solution on mesoporous birnessite. J. Taiwan Inst. Chem. Eng. 2017, 77, 168–176. [Google Scholar] [CrossRef]
  48. Abu Elella, M.H.; Sabaa, M.W.; ElHafeez, E.A.; Mohamed, R.R. Crystal violet dye removal using crosslinked grafted xanthan gum. Int. J. Biol. Macromol. 2019, 137, 1086–1101. [Google Scholar] [CrossRef] [PubMed]
  49. Zhai, Q.-Z. Studies of adsorption of crystal violet from aqueous solution by nano mesocellular foam silica: Process equilibrium, kinetic, isotherm, and thermodynamic studies. Water Sci. Technol. 2020, 81. [Google Scholar] [CrossRef]
  50. Sabna, V.; Thampi, S.G.; Chandrakaran, S. Adsorption of crystal violet onto functionalized multi-walled carbon nanotubes: Equilibrium and kinetic studies. Ecotoxicol. Environ. Saf. 2016, 134, 390–397. [Google Scholar] [CrossRef]
  51. Grassi, P.; Reis, C.; Drumm, F.C.; Georgin, J.; Tonato, D.; Escudero, L.B.; Kuhn, R.; Jahn, S.L.; Dotto, G.L. Biosorption of crystal violet dye using inactive biomass of the fungus Diaporthe schini. Water Sci. Technol. 2019, 79, 709–717. [Google Scholar] [CrossRef]
  52. Patil, S.R.; Sutar, S.S.; Jadhav, J.P. Sorption of crystal violet from aqueous solution using live roots of Eichhornia crassipes: Kinetic, isotherm, phyto and cyto-genotoxicity studies. Environ. Technol. Innov. 2020, 18, 100648. [Google Scholar] [CrossRef]
  53. Mitrogiannis, D.; Markou, G.; Çelekli, A.; Bozkurt, H. Biosorption of methylene blue onto Arthrospira platensis biomass: Kinetic, equilibrium and thermodynamic studies. J. Environ. Chem. Eng. 2015, 3, 670–680. [Google Scholar] [CrossRef]
  54. Mohamed, R.R.; Abu Elella, M.H.; Sabaa, M.W.; Saad, G.R. Synthesis of an efficient adsorbent hydrogel based on biodegradable polymers for removing crystal violet dye from aqueous solution. Cellulose 2018, 25, 6513–6529. [Google Scholar] [CrossRef]
  55. Dotto, G.L.; Salau, N.P.G.; Piccin, J.S.; Cadaval, T.R.S.; de Pinto, L.A.A. Adsorption Kinetics in Liquid Phase: Modeling for Discontinuous and Continuous Systems; Springer: Berlin, Germany, 2017; pp. 52–76. [Google Scholar]
  56. Fabryanty, R.; Valencia, C.; Soetaredjo, F.E.; Putro, J.; Santoso, S.P.; Kurniawan, A.; Ju, Y.-H.; Ismadji, S. Removal of crystal violet dye by adsorption using bentonite – alginate composite. J. Environ. Chem. Eng. 2017, 5, 5677–5687. [Google Scholar] [CrossRef] [Green Version]
  57. Piccin, J.S.; Cadaval, T.R.S.; de Pinto, L.A.A.; Dotto, G.L. Adsorption Isotherms in Liquid Phase: Experimental, Modeling and Interpretations; Springer: Berlin, Germany, 2017; pp. 19–51. [Google Scholar]
  58. Hossain, M.A.; al-Hassan, M.T. Kinetic and thermodynamic studies of the adsorption of crystal violet onto used black tea leaves. Orbital: Electron. J. Chem. 2013, 5, 148–156. [Google Scholar] [CrossRef]
  59. Laskar, N.; Kumar, U. Adsorption of Crystal Violet from Wastewater by Modified Bambusa Tulda. KSCE J. Civ. Eng. 2017, 22, 2755–2763. [Google Scholar] [CrossRef]
  60. Puri, C.; Sumana, G. Highly effective adsorption of crystal violet dye from contaminated water using graphene oxide inter-calated montmorillonite nanocomposite. Appl. Clay Sci. 2018, 166, 102–112. [Google Scholar] [CrossRef]
  61. Liu, T.; Li, Y.; Du, Q.; Sun, J.; Jiao, Y.; Yang, G.; Wang, Z.; Xia, Y.; Zhang, W.; Wang, K.; et al. Adsorption of methylene blue from aqueous solution by graphene. Colloids Surfaces B: Biointerfaces 2012, 90, 197–203. [Google Scholar] [CrossRef]
  62. Mahmoud, D.K.; Salleh, M.A.M.; Karim, W.A.W.A.; Idris, A.; Abidin, Z.Z. Batch adsorption of basic dye using acid treated kenaf fibre char: Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 2012, 181–182, 449–457. [Google Scholar] [CrossRef]
  63. Jiang, Z.; Hu, D. Molecular mechanism of anionic dyes adsorption on cationized rice husk cellulose from agricultural wastes. J. Mol. Liq. 2018, 276, 105–114. [Google Scholar] [CrossRef]
Figure 1. SEM image of adsorbent surface before (a) and after (b) crystal violet adsorption [adsorption conditions: pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1].
Figure 1. SEM image of adsorbent surface before (a) and after (b) crystal violet adsorption [adsorption conditions: pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1].
Materials 14 05861 g001
Figure 2. FTIR spectrum of Bathurst burr powder before and after crystal violet adsorption. [adsorption conditions: pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1].
Figure 2. FTIR spectrum of Bathurst burr powder before and after crystal violet adsorption. [adsorption conditions: pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1].
Materials 14 05861 g002
Figure 3. Effect of pH (a) and ionic strength (b) on adsorption capacity for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K].
Figure 3. Effect of pH (a) and ionic strength (b) on adsorption capacity for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K].
Materials 14 05861 g003
Figure 4. Effect of adsorbent dose (a) and initial dye concentration (b) on adsorption capacity and removal efficiency for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K; ionic strength: 0 mol L−1].
Figure 4. Effect of adsorbent dose (a) and initial dye concentration (b) on adsorption capacity and removal efficiency for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): pH 7; initial dye concentration: 100 mg·L−1; contact time: 30 min; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, contact time: 30 min; adsorbent dose: 2 g·L−1; temperature: 294 K; ionic strength: 0 mol L−1].
Materials 14 05861 g004
Figure 5. Effect of (a) contact time and (b) temperature on adsorption capacity for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): pH 7; initial dye concentration: 100 mg·L−1; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; ionic strength: 0 mol L−1].
Figure 5. Effect of (a) contact time and (b) temperature on adsorption capacity for the crystal violet adsorption onto Bathurst burr powder [adsorption conditions: (a): pH 7; initial dye concentration: 100 mg·L−1; adsorbent dose: 2 g·L−1; temperature: 294 K, ionic strength: 0 mol L−1 (b): pH 7, initial dye concentration: 100 mg·L−1; contact time: 30 min; adsorbent dose: 2 g·L−1; ionic strength: 0 mol L−1].
Materials 14 05861 g005
Figure 6. Comparison of experimental and predicted dye removal efficiency.
Figure 6. Comparison of experimental and predicted dye removal efficiency.
Materials 14 05861 g006
Table 1. Controllable factors and their levels.
Table 1. Controllable factors and their levels.
FactorLevel 1Level 2Level 3
pH2612
Time (min)52040
Adsorbent dose (mg·L−1)0.52.03.0
Initial dye concentration (mg·L−1)50150250
Temperature (K)285294311
Ionic strength (mol L−1)00.100.25
Table 2. The FTIR bands that were assigned to different cellulose and hemicellulose.
Table 2. The FTIR bands that were assigned to different cellulose and hemicellulose.
FTIR BandsAssignmentReference
3448 cm−1O–H stretching[31,36,38]
2340 cm−1O–H bending of adsorbed water[32]
1630 cm−1O–H bending of adsorbed water[33]
1368 cm−1C–H bending[34,36,37]
1000 cm−1C–O stretching[35]
Table 3. Adsorption isotherms model constants and the corresponding error functions.
Table 3. Adsorption isotherms model constants and the corresponding error functions.
Isotherm Model ParametersValue
Langmuir non-linearKL (L mg−1)0.024 ± 0.001
qmax (mg·g−1)164.1 ± 4.23
R20.9967
χ20.222
SSE11.47
ARE (%)2.57
Freundlich non-linearKf (mg·g−1)8.62 ± 1.47
1/n0.60 ± 0.03
R20.9828
χ21.249
SSE62.12
ARE (%)6.94
Table 4. Maximum adsorption capacities values of similar adsorbents used for the crystal violet dye adsorption.
Table 4. Maximum adsorption capacities values of similar adsorbents used for the crystal violet dye adsorption.
AdsorbentMaximum Adsorption
Capacity (mg·g−1)
Reference
Water hyacinth root powder322.58[20]
Used black tea leaves200.00[58]
Bathurst burr powder164.10This study
Moringa oleifera pod husk156.25[24]
Breadfruit skin145.80[8]
Papaya seeds85.99[19]
Pará chestnut husk83.60[21]
Pineapple leaf powder78.22[18]
Wheat bran69.15[13]
Laminaria japonica66.64[13]
Coir pith65.53[15]
Eragrostis plana nees60.10[7]
Jackfruit leaf powder43.39[17]
Rice bran41.68[13]
Sawdust37.83[15]
Date palm leaves powder37.73[22]
Peel of Cucumis sativa fruit34.24[16]
Coniferous pinus bark powder32.78[14]
Cedar cones13.64[11]
Almond shells12.20[25]
Sugarcane fiber10.44[15]
Corn stalk9.64[23]
Calotropis procera leaf4.14[12]
Table 5. Kinetic model constants and the corresponding error functions.
Table 5. Kinetic model constants and the corresponding error functions.
Kinetic Model ParametersValue
Pseudo-first order non-lineark1 (min−1)0.371 ± 0.021
qe,calc (mg·g−1)41.40 ± 0.62
R20.9960
χ20.138
SSE5.56
ARE (%)16.09
Pseudo-second order non-lineark2 (min−1)0.019 ± 0.008
qe,calc (mg·g−1)43.85 ± 0.21
R20.9999
χ20.007
SSE0.14
ARE (%)14.59
Table 6. Thermodynamic parameters for the crystal violet adsorption onto Bathurst burr powder.
Table 6. Thermodynamic parameters for the crystal violet adsorption onto Bathurst burr powder.
ΔG (kJ·mol−1)ΔH (kJ·mol−1)ΔS (J mol−1·K−1)
285 K294 K311 K
−21.62−22.51−24.351.0112.66
Table 7. Response table for signal-to-noise S/N ratios (larger is better) and contribution percentage of controllable factor influence on crystal violet removal process.
Table 7. Response table for signal-to-noise S/N ratios (larger is better) and contribution percentage of controllable factor influence on crystal violet removal process.
LevelpHIonic StrengthAdsorbent DoseInitial Dye ConcentrationTimeTemperature
127.4033.76 *29.1432.97 *31.7532.40
235.0432.3634.3432.9633.0532.66
335.70 *32.0134.65 *32.1933.33 *33.07 *
Delta8.301.745.500.781.590.68
Rank132546
Contribution (%)65.002.6029.230.622.190.36
* The maximum mean S/N ratio indicative of optimum condition.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mosoarca, G.; Vancea, C.; Popa, S.; Boran, S. Bathurst Burr (Xanthium spinosum) Powder—A New Natural Effective Adsorbent for Crystal Violet Dye Removal from Synthetic Wastewaters. Materials 2021, 14, 5861. https://doi.org/10.3390/ma14195861

AMA Style

Mosoarca G, Vancea C, Popa S, Boran S. Bathurst Burr (Xanthium spinosum) Powder—A New Natural Effective Adsorbent for Crystal Violet Dye Removal from Synthetic Wastewaters. Materials. 2021; 14(19):5861. https://doi.org/10.3390/ma14195861

Chicago/Turabian Style

Mosoarca, Giannin, Cosmin Vancea, Simona Popa, and Sorina Boran. 2021. "Bathurst Burr (Xanthium spinosum) Powder—A New Natural Effective Adsorbent for Crystal Violet Dye Removal from Synthetic Wastewaters" Materials 14, no. 19: 5861. https://doi.org/10.3390/ma14195861

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop