In recent decades, there has been extensive use of synthetic dyes in the dye-based industries, particularly the textile sector. Therefore, the aim of this study was to evaluate the adsorption of Cibacron Blue (CB) onto untreated pea pods (UPPs) in a batch system. The effects of the initial CB concentration (10–20 mg/L), solution pH (2–12), adsorbent dose (0–4 g/L), particle size (50–500 μm), and temperature (295–318 K) on the CB adsorption were investigated in batch configuration to determine the optimum conditions. Analyses of UPPs were characterized by Fourier Transform Infrared (FTIR) spectroscopy, Brunauer–Emmett–Teller (BET), and point-of-zero charge (pHpzc = 5.6). Under optimized conditions (pH: 2.5, particle size: 50 μm, time: 40 min, adsorbent dose: 2.5 g/L, and agitation speed: 250 rpm), up to 30.30 mg/g at 25 °C is removed from the solution. The adsorptions kinetics obey rather a pseudo-second-order kinetic model with a determination coefficient of R2 = 0.999. The adsorption isotherms have been used for the determination of thermodynamic parameters, i.e. the negative free energy ΔG° (−4.33 to 0.783 kJ/mol), negative enthalpy change ΔH° (−54.63 kJ/mol), and entropy (ΔS° = −0.1705 kJ/mol.K) indicate that the CB adsorption onto UPPs is spontaneous and exothermic in nature.

  • In the present study, the adsorption of CB by untreated pea pods was studied and the maximum adsorption capacity deduced was 30.30 mg/g at 25 °C.

  • The adsorption kinetics in perfect correlation with the pseudo-second-order kinetic model with a determination coefficient (R2) of 0.999 was obtained.

  • The thermodynamic parameters determined indicate that the adsorption of CB is spontaneous and exothermic in nature.

With the rapid development of industry, organic dyes are widely used in fields such as textiles, pharmaceuticals, leather, and cosmetics and it has been stated that at least 15% of the total dye is wasted each year. Substantial amounts of synthetic dyes, extremely resistant to biodegradation, have accumulated in the environment. Dyes enter the human body through food and skin contact, so reducing their application to values below the limit is also an important ecological and medical problem. Several industrial sectors stand out in the environmental scenario as major pollutants, mainly in the textile field, due to the volume of effluents generated by synthetic dyes that do not bind completely to the fiber during the dyeing process (Inyinbor et al. 2016). The widespread use of synthetic dyes has caused serious problems for human health and aquatic life due to their mutagenic, carcinogenic, and toxic properties, and colored wastewater is a direct result of their production. Indeed, many dyes and their degradation products are toxic to living organisms (Abbas 2020, 2021a, 2021b, 2021c, 2022a; Abbas & Trari 2023a). The discharge of industrial effluents generated by many sectors, in particular paints, plastics, paper, leather, and textiles, is becoming a major concern due to their negative effects on the environment. The availability of fresh water is an important requirement for growth and development. It is not only a basic human convenience, but also a prime requirement for development. Various water-intensive activities, such as irrigation, sanitation, and hydropower generation, are dependent on the availability of freshwater resources (Wallace 2000). Developed countries are vulnerable to climate change and critical constraints to food security depend on rainfall. The impacts of climate variability are manifested mainly through changes in temperature and precipitation, which are expected to reduce water availability (Vörösmarty et al. 2010). Among the direct consequences of human-induced climate change, one can cite the following: (i) rising temperatures, (ii) sea level rise, (iii) increasing periods of aridity and drought, and(iv) melting glaciers. More details about the climate changes and their consequences and are given elsewhere (Vörösmarty et al. 2010). Dyes are characterized by high molecular weight and complex chemical structures make organic dyes difficult to degrade under natural conditions. Moreover, most organic dyes are mutagenic, teratogenic, and carcinogenic biological toxins, as well as highly soluble and color saturated, which seriously jeopardizes the safety of biological species and global ecosystems. Cibacron Blue (CB) is one such dye, commonly used in dyeing and textiles, and it is poorly biodegradable. Thus, due to the harmful effects, wastewater containing CB must be treated at the source. Therefore, proper treatment of industrial waters that release lead into aquatic and terrestrial environments is of paramount importance (Singanan et al. 2005). Some processes, including mechanical (e.g., filtration and reverse osmosis), physical (e.g., adsorption, extraction, and flocculation), chemical (e.g., precipitation, oxidation, ion exchange, and sonolysis), and thermal (e.g., evaporation and distillation) methods have been used to remediate the contaminated water by dyes and other pollutants (Abbas & Trari 2020a, 2020b; Abbas et al. 2020; Abbas 2022b). Among the treatment strategies, adsorption is considered the most universal water treatment technology and preferable method for removing pollutants from wastewater; this method is low in cost, with simplicity of design, high removal efficiency, ease of operation, and availability. Agricultural waste that is evaluated for heavy metal and dye removal has advantages as an adsorbent: readily available and existing in abundance, it is cost-effective, renewable, requires less processing time, offers suitable adsorption capability, and can be easily regenerated. Agricultural waste can be modified by treating it with different chemical agents, e.g., alkalis, acids, organic compounds, etc., or thermally – this modification could have beneficial effects on chemical/physical properties including increasing surface area (SA), improving pore structure, adding a functional group, for instance, amino (–NH2), carboxyl (–COOH), and hydroxyl (–OH) on their surface. Modified adsorbents exhibit adsorption capacity, are more selective and sensitive. Many agricultural wastes such as cotton stalk, olive stones, coconut shell, rice husk, orange peel, pistachio shell, wheat bran, walnut shell, pomegranate peel, natural clay, Lagerstroemia speciosa seed, and biomaterials are used in adsorption applications (Abbas 2022b, c; Abbas & Trari 2023b, 2023c; Merrad et al. 2023; Das et al. 2024). Today, global challenges focus on the following: reducing the environmental impact caused by water pollution, reducing industrial waste that can cause an imbalance in the ecosystem, and raise public awareness of the consequences of global warming. The main objective of this study is to recover natural waste in the water treatment. The study aims to examine the adsorption capacity of untreated pea pods prepared for the adsorption of CB. UPP was characterized by physicochemical methods and the effect of operational parameters, namely pH, contact time, adsorbent dose, particle size, CB concentration, and temperature, were undertaken. Similarly, the experimental data were subjected to equilibrium kinetics and linear regression modeling and validated using a statistical error model. The novelty and ambitions of this study are focused on the following points: (1) reduce the environmental impact caused by water pollution; (2) propose solutions to reduce industrial waste that can cause an imbalance in the ecosystem;(3) raise public awareness of the consequences of global warming; (4) environmental protection means protecting humanity for life on earth; (5) comply with the rules relating to the discharge of industrial waste; (6) show that the prepared adsorbent can constitute a new potential candidate in water treatment comparable to expensive , activated carbon; and (7) finally tests on an industrial scale constitute the logical continuation of this study.

Materials

CB (99%, C.I. number: 621526) is an anionic dye, supplied by the textile industry and used without any further purification; its structure and chemical characteristics are presented in Table 1. Colored solutions were prepared by dissolving the required CB in distilled water to produce a stock solution of 1 g/L. Adsorption studies for the evaluation of adsorbent for the CB removal from aqueous solutions were carried out in a series of 100-mL flasks using a batch contact adsorption method. UV–Visible spectrometer Shimadzu/model UV-1601 PC was used to measure absorbance of CB solution (λmax = 625 nm; Figure 1). The CB concentration was determined from its UV–visible absorbance using a calibration graph.
Table 1

General characteristics of CB

PropertiesStructural formula
Molecular weight 882.17 g/mol  
Wavenumber (λmax625 nm 
Name Reactive Blue 49 
 Cibacron Bleu 3G-A 
Solubility in water 20 mg/mL at T= 25 °C 
C.I. number 621526 
CAS Registry Number 12236-92-9 
Company Sigma–Aldrich 
Purity 99% 
PropertiesStructural formula
Molecular weight 882.17 g/mol  
Wavenumber (λmax625 nm 
Name Reactive Blue 49 
 Cibacron Bleu 3G-A 
Solubility in water 20 mg/mL at T= 25 °C 
C.I. number 621526 
CAS Registry Number 12236-92-9 
Company Sigma–Aldrich 
Purity 99% 
Figure 1

Spectrum of Reactive Blue in aqueous solutions.

Figure 1

Spectrum of Reactive Blue in aqueous solutions.

Close modal
The pea pods (Pisum sativum) used in this work were obtained after shelling the vegetables (Figure 2). The preparation of pea pod powder involves four main steps: drying, grinding, washing, and sieving (Figure 3). The drying of the material was carried out using solar energy and also in an oven, between 50 and 60 °C (12 h). In order to prevent a possible alteration of the physicochemical properties of the materials, for the adsorption tests, the pea pods were crushed to obtain homogeneous materials and to increase their adsorption capacity. Indeed, the use of powdered materials is recommended by many authors. Washing was carried out by contacting a mass of material obtained after grinding with a quantity of distilled water (10 g/L). The suspension obtained was stirred using a propeller stirrer for 24 h at a speed of 110 rpm. The material was dried again at a moderate temperature between 50 and 60 °C, until a constant mass was obtained. The powder obtained is sieved using a sieve of different particle sizes in the range (50–500 μm).
Figure 2

Natural state Pisum sativum L. of the precursor used as an adsorbent (UPP).

Figure 2

Natural state Pisum sativum L. of the precursor used as an adsorbent (UPP).

Close modal
Figure 3

Steps of adsorbent preparation (UPP).

Figure 3

Steps of adsorbent preparation (UPP).

Close modal

Experimental characterizations

The SA of the sample clay was determined by the Brunauer–Emmett–Teller (BET)method using a AsiQuin, Automated Gas Sorption Analyzer Quantachrome Instrument Version 2.02. The specific SA and pore structure of the activated carbons were characterized by N2 adsorption–desorption isotherms at −196 °C using the ASAP 2010 Micromeritics equipment.

Fourier Transform Infrared (FTIR) spectroscopy was used to identify the characteristic functional groups of commercial clay. Approximately 20mg of (UPP) was mixed with 80mg of dry KBr of spectroscopy quality and pressed under a pressure of 4,500 psi to form a thin disc. Then, the FTIR spectra were plotted with a Perkin Elmer 2000 infrared spectrometer in the range (4,000–400cm−1) 16 times to increase the signal-to-noise ratio.

To observe the surface structure of the adsorbent (UPP) before adsorption of CB, scanning electron micrographs (SEMs) are produced with different resolutions using a Scanning Electron Microscope (JOEL-5910).

The concentration of CB content in the supernatant is determined using UV–visible spectrophotometer, and the pH of the solutions was accurately measured using a microprocessor-based pH meter of the HANNA HI 8521 type. The instrument was calibrated with commercial buffers of pH 4, pH 7, and pH 10.

The zero-point-charge pH(zpc) of UPP, i.e. the pH for which the surface charge is zero, is determined using a procedure similar to that described elsewhere (Merrad et al. 2023). Approximately 20 mL of KNO3 solutions (0.01 M) are placed in closed conical flasks; the pH of each solution is adjusted between 2 and 14 by adding HCl or NaOH solution. Then, 0.1 g of UPP is added and the final pH is measured after 24 h under agitation at room temperature. The zero-point-charge pH(zpc) is the final pH versus initial pH crosses the line at final pH = initial pH.

Adsorption experiments

The effects of the initial CB concentration (10–20 mg/L), solution pH (2–12), adsorbent dose (0–4 g/L), particle size (50–500 μm), and temperature (295–318 K) on the CB adsorption were investigated in batch configuration for variable specific periods (0–50 min). The CB solutions were made up by dissolving the accurate amount of CB (99%) in distilled water, used as the stock solution and diluted to the required concentration; pH was adjusted with HCl (0.1 mol/L) or NaOH (0.1 mol/L). For the kinetic studies, desired quantities of UPPs were mixed with 10 mL of CB solutions in Erlenmeyer flasks and placed on a rotary shaker at 300 rpm. The aliquots were withdrawn at regular times and subjected to centrifugation at 3,000 rpm (10 min). The adsorbed quantity qt (mg/g) and the percentage of CB elimination Rt (%) by UPPs were calculated from the following equations:
(1)
(2)

C0 is the CB initial concentration and Ct is the CB concentrations (mg/L) at time (t), V is the volume of solution (L), and m is the mass of UPP (g).

Analyses of the surface chemistry of UPPs

The FTIR spectrum (Figure 4) of UPP displays a number of adsorption peaks, indicating that many functional groups of the adsorbent are involved in the adsorption, and the indexing of these bands by the international spectroscopy table shows that the band 3,447 cm−1 is assigned to the elongation of the group O–H, with the absorption bands centered at 2,924 cm−1cm relating to the elongation of the C–H group. The band at 1,636 cm−1 is attributed to the bonding vibration of C = O carbonyl groups relating to the oxygenated groups of lignins (Liu et al. 2009), and the band centered at 1,427 cm−1 corresponds to the carboxylic acid of pectins and hemicelluloses, as well as to the vibrations of carboxylates. The band at 1,033 cm−1 is also ascribed to characteristic of the elongation in the plane of the connection.
Figure 4

IR spectrum of the prepared adsorbent (UPP).

Figure 4

IR spectrum of the prepared adsorbent (UPP).

Close modal

BET and SA analysis

Gas sorption measurements were used to analyze the UPP pore parameter. More specifically, nitrogen adsorption/desorption Figure 5 at the surface of solid powders is the most widespread method for determining the SA of UPP. Nitrogen adsorption and desorption curves N2 and the BET, SA finding shows to the structure of the mean pore diameter and the mean SA (66.66 m2/g) of UPP. The shape of the adsorption–desorption isotherms obtained on the adsorbent is similar to type II of the IUPAC classification due to the absence of the saturation plateau at high relative pressures that characterizes the shape of type IV isotherms. The hysteresis loop observed from these isotherms appears to be of the H type, which is observed with solids composed of aggregates of flat particles or containing slotted pores.
Figure 5

Adsorption and desorption isotherms of nitrogen at 77 K by UPP.

Figure 5

Adsorption and desorption isotherms of nitrogen at 77 K by UPP.

Close modal

Scanning electron microscopy

The SEMs of UPP before and after adsorption are shown in Figure 6; the prepared UPP presents a microporous structure with different pore diameters, a rough surface, and many protrusions. After adsorption, the UPP surface became smoother where the roughness was considerably reduced with less visible pores, indicating a clear adsorption on both the surface and within pores. The images also reveal that the exterior surfaces of the UPP are filled with more or less homogeneous cavities of different sizes and shapes. These cavities differ from one carbon to another during the reaction with the activating agent. These cavities are the external pores that represent the main channels for accessing the internal surface of the activated carbon. The highest magnification shows that the absorbant surface contains a considerable number of pores with a high probability that CB molecules are adsorbed.
Figure 6

Microscopic observation of the adsorbent (a, b) before and (c, d) after CB adsorption.

Figure 6

Microscopic observation of the adsorbent (a, b) before and (c, d) after CB adsorption.

Close modal

Studies of the effect of process variables

Effect of UPP size

In the first stage of the batch adsorption experiments, the effect of the particle sizes (50–500 μm) on the CB adsorption onto UPP is examined. Significant variations in the uptake capacity and removal efficiency are observed for different particles sizes. Figure 7 indicates that the best performance is obtained with smaller sizes (50 μm). In general, smaller particles provide large SA, thus resulting in high uptake capacity and removal efficiency. Therefore, the class (50 μm) was subsequently used in all adsorption experiments.
Figure 7

Influence of the particle size on the adsorption capacity (C0 = 20 mg/L, pH = 2.5, V= 25 mL, T= 25 °C, adsorbent dose: 1 g/L, agitation speed: 250 trs/min, time = 50 min).

Figure 7

Influence of the particle size on the adsorption capacity (C0 = 20 mg/L, pH = 2.5, V= 25 mL, T= 25 °C, adsorbent dose: 1 g/L, agitation speed: 250 trs/min, time = 50 min).

Close modal

Effect of contact time and initial concentration of CB

The adsorption capability of UPP was assessed in relation to the initial CB concentration at various contact times, as shown in Figure 8. The level of adsorption rose along with the initial CB concentration, which may be related to the greater concentration gradient as it acts as the driving force for CB species to adsorb onto the active UPP surface sites (Jawad & Abdulhameed 2020). The adsorption capacity of CB increases over time and reaches a maximum after 40 min of contact time and thereafter, with a constant value indicating that no more CB ions are further removed from the solution.
Figure 8

Evolution of the CB adsorption at different concentrations on UPP as a function of time (C0 = 10 and 20 mg/L, V= 25 mL, adsorbent dose: 2.5 g/L, t = 60 min, agitation speed: 250 trs/min T= 25 °C, particle size: 50 μm, time: 60 min).

Figure 8

Evolution of the CB adsorption at different concentrations on UPP as a function of time (C0 = 10 and 20 mg/L, V= 25 mL, adsorbent dose: 2.5 g/L, t = 60 min, agitation speed: 250 trs/min T= 25 °C, particle size: 50 μm, time: 60 min).

Close modal

The equilibrium times averages 40 min but for practical reasons the adsorption experiments are run up to 50 min. With raising the initial CB concentration (C0) from 10 to 20 mg/L, the adsorbed quantity of CB onto UPP rose from 3.86 to 8.20 mg/g. The CB ions are adsorbed initially on the external SA of UPP, which makes the adsorption rate easy and fast. When the adsorption of the external surface reaches saturation, the CB ions entered into the pores and absorb on the internal surface of the particles and such a phenomenon takes relatively longer contact time. Similar behavior was reported in the literature (Hammeed et al. 2007). The rapid adsorption kinetics recorded during the first minutes can be explained by the availability of a large number of active sites on the surface of the adsorbent material at the beginning of the adsorption process.

The decrease in the rate of adsorption over time is due to the occupation of the still vacant sites and the appearance of repulsion forces between the lead ions adsorbed by the phosphates and those that are in solution. Finally, saturation is reached after 40 min. This may be attributed to an increase of the driving force due to the concentration gradient with increasing C0 in order to overcome the mass transfer resistance of CB between the aqueous and solid phases. Therefore, a higher initial CB concentration C0 increases the adsorption capacity.

Effect of adsorbent dosage

The influence of the adsorbent dose on the amount of CB adsorbed was studied by bringing the dye solution into contact at an initial concentration of 20 mg/L with the dose of adsorbent, which varies from 1 to 4 g/L. The results obtained are shown in Figure 9. The preceding curve clearly reveals that the quantity of adsorption increases with the increase in the dose of adsorbent in the range 1–3 g/L. This can be explained by the fact that many effective and active sites are used at a higher adsorbent dose due to the aggregation and overlapping of adsorbent particles in the solution, resulting in a decrease in SA of absorption accessible by CB molecules. In the range 3–4 g/L, a decrease in the amount of adsorption is observed, which can be explained by the fact that there is a meeting between the particles of the adsorbent of the same charge. This causes electrostatic repulsions and unfavorable adsorption, the maximum of which is obtained for a dose of 3 g/L, which will be retained for the rest of the experiments.
Figure 9

Evolution of the adsorption of CB onto UPP as a function of the adsorbent dose (C0 = 20 mg/L, pH = 2.5, V= 25 mL, agitation speed: 250 trs/min, T= 25 °C, particle size: 50 μm, time: 40 min).

Figure 9

Evolution of the adsorption of CB onto UPP as a function of the adsorbent dose (C0 = 20 mg/L, pH = 2.5, V= 25 mL, agitation speed: 250 trs/min, T= 25 °C, particle size: 50 μm, time: 40 min).

Close modal

Effect of pH

The pH is one of the important parameters that influence the adsorption of dyes. This is in addition to its impact on the solubility of dyes in solution; H+ ions can replace positive ions present in the active sites and affect the degree of ionization of the adsorbate. Indeed, pH not only has an influence on the surface charge of UPP but also affects the structure and functional groups of CB. Its effect on the adsorption can be explained by the zero-point-charge pH(pzc), where the surface functions of the material have a significant influence on the adsorption performance (Deng et al. 2011). The basic or acidic nature of the adsorbent surface governs the retention capacity of the pollutant. However, the character and chemical properties of the adsorbent are directly related to the nature of the functional groups on its surface. The surface charge of the adsorbent, resulting from the acid–base balance, depends on both the ionic strength of the solution and the pH. This charge can be positive or negative depending on environmental conditions. Therefore, an important characteristic of the surface is the determination of pH(pzc) (=5.6), as seen in Figure 10, by the drift method, which defines the pH for which the surface charge, linked to the exchange of protons, is zero; pH(pzc) characterizes the acidity or alkalinity of the adsorbent surface UPP. The surface charge is positive below pH(pzc), where the oxygen groups are in cationic form, which constitutes a favorable medium for the adsorption of CB, an anionic dye. This is due to the attractive electrostatic forces that increase the CB elimination efficiency. Conversely, above pH(pzc) both the adsorbent surface and CB are negatively charged, thus causing repulsive electrostatic forces that reduce the CB removal efficiency.
Figure 10

Evolution of the CB adsorption onto UPP as a function of pH (C0 = 20 mg/L, pH variable, V= 25 mL, adsorbent dose: 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, T= 25 °C, and particle size: 50 μm).

Figure 10

Evolution of the CB adsorption onto UPP as a function of pH (C0 = 20 mg/L, pH variable, V= 25 mL, adsorbent dose: 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, T= 25 °C, and particle size: 50 μm).

Close modal

Adsorption isotherms modeling

The aim of this section is to understand the interaction between dye and adsorbent through the validity of the models and to find parameters allowing comparison, interpretation, and prediction of the adsorption data of UPP. To assess the effectiveness of the adsorbent, different equations exist and are applied at optimal conditions (Figure 11).
Figure 11

Isotherms of different models at optimum conditions (C0 = variable, V= 2mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5; T= 25 °C, particle size: 50 μm).

Figure 11

Isotherms of different models at optimum conditions (C0 = variable, V= 2mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5; T= 25 °C, particle size: 50 μm).

Close modal
The Langmuir model (Langmuir 1918) postulates the occurrence of monolayer adsorption onto fixed number of localized sites on an adsorbent surface. The model further assumes that a given adsorbent surface is composed of sites homogeneously equivalent in their enthalpies but with no subsequent movement of adsorbed molecules in the surface plane, and no interactions between neighboring adsorbate molecules; the non-linear expression is as follows:
(3)
qmax is the monolayer adsorption capacity (mg/g) and KL is the constant (l/mg) related to the free adsorption energy and is used to determine the dimensionless separation factor RL. It indicates the favorability of the adsorption process and is given by Equation (4):
(4)

RL indicates the type of isotherm: irreversible (RL = 0), favorable (0 < RL < 1), linear (RL = 1), or unfavorable (RL > 1). In this contribution, the RL values are smaller than 1, thus confirming that the adsorption is favorable in both cases as well as the applicability of the Langmuir isotherm.

Freundlich's model (Freundlich 1906) is based on the formation of unlimited multilayers of adsorbed species, with an infinite surface coverage predicted on a heterogeneous surface. The enthalpies of the adsorbent surface sites follow a logarithmic distribution, where the higher energy sites with greater affinity for the adsorbate are occupied first, followed by the lower energy sites. The sorption process is summed across sites, and the non-linear expression of the Freundlich model is given as follows:
(5)
KF (l/g) and n are the Freundlich constants, related, respectively, to the capacity of adsorption and favorability of adsorption; the plot lnqe versus lnCe enables us to extract the constant KF and n. The constant n indicates the favorability of the adsorption process. When value is between 2 and 10, favorable adsorption is expected, while an ‘n value’ less than unity indicates poor sorption characteristics.
Temkin's model (Temkin & Pyzhev 1940) postulates the heterogeneity of an adsorbent surface, whose adsorption energy distribution is linear; the non-linear form is given as follows:
(6)

where is the Temkin adsorption energy change and is the maximum adsorption capacity. T (K) is the absolute temperature and R is the universal gas constant. The adsorption data are analyzed according to Equation (8) and the linear plot qe versus lnCe permits to calculate the constants AT and BT.

The Elovich model (Cheung et al. 2000) is based on the principle of the kinetics, assuming that the number of adsorption sites augments exponentially with the adsorption, thus implying a multilayer adsorption described by the following:
(7)
where KE (l/mg) is the Elovich constant at equilibrium, qmax (mg/g) is the the maximum adsorption capacity, qe (mg/g) is the adsorption capacity at equilibrium, and Ce (g/l) is the concentration of the adsorbate at equilibrium. Both the constants KE and qe are calculated from the plot of ln(qe/Ce) versus qe. The constants of all models deduced after modeling are grouped in Table 2.
Table 2

Parameters of the adsorption isotherms for CB dye onto UPP

LangmuirFreundlichTemkinElovich
KLqmax 0.0308 L/g
30.30 mg/g 
1/n: 0.5157
n: 1.942
KF: 2.162 mg/g 
BT: 6.258
AT: 0.348 L/g
ΔQ: 12.019 J/mol 
KE: 0.0995 L/g
qmax: 12.98 mg/g 
R2 RSS 0.997
0.069 
0.943
0.153 
0.995
1.569 
0.935
0.156 
LangmuirFreundlichTemkinElovich
KLqmax 0.0308 L/g
30.30 mg/g 
1/n: 0.5157
n: 1.942
KF: 2.162 mg/g 
BT: 6.258
AT: 0.348 L/g
ΔQ: 12.019 J/mol 
KE: 0.0995 L/g
qmax: 12.98 mg/g 
R2 RSS 0.997
0.069 
0.943
0.153 
0.995
1.569 
0.935
0.156 

RSS, residual sum of squares; R2, determination coefficient; ΔQ, Temkin energy.

Adsorption kinetic modeling

The kinetics of CB adsorption is crucial to determining the operating conditions that are optimized for a full-scale batch process. It gives the uptake rate of the adsorbate, controls the residual time of the global process, and predicts both the adsorption rate and designing of adsorption. Here also, different models were proposed to study the behavior of adsorbents and to propose the mechanisms controlling the adsorption. The experimental data of CB adsorption are examined using the pseudo-first- and pseudo-second-order kinetic models given by (Ho & McKay 1998; Lagergen 1898):

Pseudo-first-order model:
(8)
Pseudo-second-order model:
(9)
For the pseudo-second-order, the initial adsorption rate h (mg/g.min) is expressed by
(10)
where qt (mg/g) is the adsorbed amount of CB on UPP at the time t (min), K1 (min−1) and K2 (g/mg.min) are the pseudo-first-order and pseudo-second-order rate constants, respectively. The slope and intercept of the plots ln(qe–qt) versus time (t) are used to calculate the first-order rate constants K1 and qe while the plot of t/qe versus t is used to determine K2 and qe, which predict the CB uptake.

For the pseudo-first-order kinetic, the experimental data deviate from linearity, as evidenced by the low values of qe and Co, suggesting the inapplicability of the model for the present system. By contrast, the determination coefficient and qe,cal from the pseudo-second-order model agree perfectly with the experimental kinetic data; the corresponding coefficients for UPP are summarized in Table 3.

Table 3

Pseudo-first-order and pseudo-second-order, Elovich, and diffusion model constants and determination coefficients for CB adsorption onto UPPs

Second-order
Pseudo-first-order
C0 (mg/L)qex (mg/g)qcal (mg/g)R2SSEK2 (g/mg.mn)qcal (mg/g)R2SSE (%)K1 (mn−1)
10
20 
3.9
7.9 
4.29
8.20 
0.998
0.999 
0.073
0.002 
0.057
0.058 
2.899
6.266 
0.932
0.596 
0.007
0.012 
0.0836
0.0116 
Elovich
Diffusion
C0 (mg/L)R2β (g/mg)α (mg/g.mn)SSEKin (mg/g.mn1/2)R2C (mn1/2)RSS
10
20 
0.991
0.992 
0.498
0.182 
2.578
14.92 
0.002
0.003 
 0.20.8
0.253 
0.994
0.956 
2.551
6.178 
0.002
0.021 
Second-order
Pseudo-first-order
C0 (mg/L)qex (mg/g)qcal (mg/g)R2SSEK2 (g/mg.mn)qcal (mg/g)R2SSE (%)K1 (mn−1)
10
20 
3.9
7.9 
4.29
8.20 
0.998
0.999 
0.073
0.002 
0.057
0.058 
2.899
6.266 
0.932
0.596 
0.007
0.012 
0.0836
0.0116 
Elovich
Diffusion
C0 (mg/L)R2β (g/mg)α (mg/g.mn)SSEKin (mg/g.mn1/2)R2C (mn1/2)RSS
10
20 
0.991
0.992 
0.498
0.182 
2.578
14.92 
0.002
0.003 
 0.20.8
0.253 
0.994
0.956 
2.551
6.178 
0.002
0.021 

The applicability of the model suggests that the adsorption CB onto UPP is based on a physical reaction (physisorption), involving an exchange of electrons between the adsorbent and the adsorbate where the CB ions are attached to the adsorbent surface by chemical bond.

The Elovich kinetic equation is related to the process and is often validated for systems where the surface of the adsorbent is heterogeneous (Chien & Clayton 1980); the linear form is given by
(11)
where α (mg/gmin) is the initial adsorption rate and β (mg/g) is the relationship between the degree of surface coverage and the activation energy involved in the chemisorption.

Intraparticle diffusion equation

Generally, a process is diffusion controlled if its dependence on the rate at which the components diffuse toward one another is the limiting step. When the adsorption rate is controlled by an intraparticle diffusion mechanism which is a physical step in the adsorption process, the plot of the curve qt (mg/g) as a function of time t1/2 is a straight line passing through the origin. The possibility of the intraparticle diffusion (Weber & Morris 1963) during the transport of adsorbate from solution to the particles’ surface was investigated by using the following intraparticle diffusion model:
(12)
where Kid is the intraparticle diffusion rate (mg/g min1/2), qt is the amount of CB adsorbed at time t, and C (mg/g) is the intercept. A plot of qt versus t1/2 enables to determine both the constant Kin and C. According to Aniagor et al. (Aniagor & Menkiti 2018), the resulting figure (not shown) presents a multi-linearity correlation, which indicates that two steps occur during the CB adsorption.

The mechanism of adsorption is complex, but the intraparticle diffusion is important in the early stages.

  • (i) The first linear portion is due to intraparticle diffusion.

  • (ii) The slopes of the linear parts are defined as rate parameters, characteristic of the adsorption rate in the region where the intraparticle diffusion occurs.

Initially and within a short time period, it is postulated that CB is transported to the adsorbent external surface through the film diffusion with a high rate.

After the surface saturation, the CB ions enter inside the adsorbent by intraparticle diffusion through the pores and internal surface diffusion until equilibrium is reached, evidenced by the second straight lines.

The constants of the different models deduced after modeling are grouped in Table 3, which depicts an intraparticle diffusion C value, and based on literature reports, such a high value (C > 1) is synonymous with the occurrence of substantive boundary layer effects within the system. Thus, intraparticle diffusion cannot be regarded as a probable sorption mechanism in the present study.

Thermodynamic parameters of adsorption

The temperature is an important parameter that considerably influences the adsorption of CB on the surface of the UPP adsorbent. Its importance is not limited only to the need to understand how it affects adsorption and desorption, but also to the use of experimental results, allowing access to useful thermodynamic information, namely, the enthalpy and the entropy. They can be evaluated from the modeling of experimental results with adequate theoretical models. The variation in the adsorption capacity of CB as a function of temperature is illustrated in Figure 12, it is noted that the increase in temperature from 25 to 50°C causes a decreasing in the efficiency of adsorption from 93 to 62%. This reveals that the rise in temperature disfavors the process of adsorption, indicating that the adsorption has an exothermic nature. The thermodynamic functions including standard Gibbs free energy (ΔG°), standard entropy (ΔS°), and standard enthalpy (ΔH°) were explored to provide a clear insight into the adsorption process of CB onto UPP regarding spontaneity and irregularity. Hence, Equations (13)–(16) are used to compute the adsorption thermodynamic properties such as ΔG°, ΔH°, and ΔS°, respectively (Jawad et al. 2022):
(13)
(14)
(15)
(16)
where R is the gas constant (8.314 J/mol K), T (K) is the temperature where Kd is the distribution coefficient, qe (mg/g) is the quantity of CB adsorbed at equilibrium, and Ce (mg/L) is the quantity of CB remaining in the solution at equilibrium. The values of ΔH° and ΔS° were attained from the plot of ln Kd against (1/T)Figure 13, which can be computed from the slope and the intercept, respectively. The thermodynamic functions for the adsorption of CB by UPP are listed in Table 4. The negative values of ΔG° confirmed that the CB adsorption was spontaneous and decreasing at higher temperatures. The CB adsorption on the UPP surface can be considered a physisorption based on the values of ΔG°. The fact that the ΔH° value is positive indicates that the CB adsorption process was endothermic in nature.
Table 4

Thermodynamic functions ΔG°, ΔS°, and ΔH° of CB adsorbed onto UPP

T (K)KEΔH° (kJ/mol)ΔS° (kJ/K.mol)ΔG° (kJ/mol)
295 5.81 − 54.63 − 0.1705 −4.33 
305 2.85 −2.63 
315 1.46 −0.923 
325 0.74 0.783 
T (K)KEΔH° (kJ/mol)ΔS° (kJ/K.mol)ΔG° (kJ/mol)
295 5.81 − 54.63 − 0.1705 −4.33 
305 2.85 −2.63 
315 1.46 −0.923 
325 0.74 0.783 
Figure 12

Influence of temperature on the adsorption capacity of CB onto UPPs (V= 25 mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5, T= variable, particle size: 50 μm).

Figure 12

Influence of temperature on the adsorption capacity of CB onto UPPs (V= 25 mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5, T= variable, particle size: 50 μm).

Close modal
Figure 13

Regression of the thermodynamic parameters of CB onto UPPs (V= 25 mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5, T= variable, particle size: 50 μm).

Figure 13

Regression of the thermodynamic parameters of CB onto UPPs (V= 25 mL, adsorbent dose 2.5 g/L, t = 40 min, agitation speed: 250 trs/min, pH: 2.5, T= variable, particle size: 50 μm).

Close modal

Adsorption mechanism

As previously mentioned, intraparticle diffusion is not the only factor limiting the CB adsorption onto UPP, and other mechanisms can be involved simultaneously in the adsorption process of this system. In order to confirm the functional groups involved in the mechanism of CB adsorption, we compared the FTIR spectra of the UPP carbon before and after the CB adsorption. The examination of the IR spectra obtained after adsorption confirms the attachment of CB to the UPP surface. In fact, we note the appearance of three absorption bands, specific to the O–H, C = C, and C–O groups. The position of these bands is displaced, compared to the native activated carbon, reflecting the existence of notable interactions between the CB molecules and the active sites of the carbon surface. The shift of the absorption band corresponding to the hydroxyl group O–H confirms the formation of a hydrogen bond between CB and the activated carbon. Hydrogen bonding has occurred between H–donor atoms and acceptor groups on the surface of activated carbon. The characteristic band of aromatic group C = C is shifted and this reflects the presence of hydrophobic interactions of π–π type (Figure 14), manifested between the π electron donor group of the aromatic ring and the acceptor group on the activated carbon surface. The characteristic C–O absorption band is also shifted with respect to that of the native activated carbon, thus attesting to the strong n–π type interactions involved between the oxygen groups of the activated carbon, acting as an electron donor, and the aromatic ring. It should be noted that the pHzpc of the adsorbent is 5.5.
Figure 14

Adsorption mechanism probable.

Figure 14

Adsorption mechanism probable.

Close modal

For pH values greater than 5.5, there is a predominance of negative charges on the surface of the adsorbent and since the CB is negatively charged, we notice repulsive electrostatic forces (two charges with the same signs), therefore the adsorption of iodine is practically impossible except for anions linked by hydrogen bonds or by Kesson forces (very weak adsorption). For pH lower than 5.5, there is a predominance of positive charges on the surface of the adsorbent, and since the CB is negatively charged, we notice attractive electrostatic forces (two charges with opposite signs), therefore the adsorption of iodine in practice is very favorable. Figure 10 clearly illustrates the predominant zones of positives and negatives charges of the adsorbent as a function of pH as well as the nature of the electrostatic forces involved in relation to the negative charge of the pollutant in the pH range studied from 2 to 8.

Performance of the adsorbent

In order to reveal the reference of the as-prepared UPP for the removal of CB, a comparative study was undertaken with the other adsorbents in the literature (Table 5) where the maximum Langmuir adsorption capacities qmax of the different adsorbents are reported. It is shown that adsorption in the present study are well classified with qmax capacity of 30.30 mg/g at 25 °C. UPP is an interesting adsorbent for basic and acidic dyes owing to its isoelectric point pH(pzc). Besides the many factors that have been optimized such as pH, temperature, stirring speed, initial dye concentration, and contact time, the regeneration of the prepared UPP is of intrinsic importance for the removal of dyes, not only to increase the reuse of the adsorbent, but also to improve its commercial viability.

Table 5

Comparison of maximum adsorption capacities for CB dye with literature data

Adsorbentsqmax (mg/g)Ref.
Agave americana fibers 8.54 Irish & Irish (2000)  
11.2 
Maize cob, Citrus peel, and rice husk powders 18.58 Saroj et al. (2015)  
Citrus waste biomass (immobilized) 80.00 Asgher (2012)  
Citrus waste biomass (raw) 135.16 – 
Citrus waste biomass (acetic acid treated) 232.56 – 
Capsicum annuum seeds (acetone treated) 96.35 Akar et al. (2011)  
Waste biomass of Aspergillus fumigatus 60.6 Wang et al. (2016)  
A. niger powder 29.96 Xiong et al. (2010)  
AL 20.83 Bouhadjra et al. (2021a)  
ALa 50.76 – 
ALc 102.04 – 
Untreated peanut shell (UPS) 30.30 Abbas & Trari (2023d)  
Activated biochar from oil palm 393.97 Jabar & Odusote (2020)  
Agricultural waste products 8.058 Habeeb & Ghawi (2023)  
Agricultural waste products (HCL) 6.514 – 
Agricultural waste products (KOH) 4.17 – 
Native potato peel powder (PP) 85 Bouhadjra et al. (2021b)  
Potato peel powder activated (PPa) 108 – 
Potato peel powder calcined (PPc) 268 – 
Untreated adsorbent 30.40 This study 
Adsorbentsqmax (mg/g)Ref.
Agave americana fibers 8.54 Irish & Irish (2000)  
11.2 
Maize cob, Citrus peel, and rice husk powders 18.58 Saroj et al. (2015)  
Citrus waste biomass (immobilized) 80.00 Asgher (2012)  
Citrus waste biomass (raw) 135.16 – 
Citrus waste biomass (acetic acid treated) 232.56 – 
Capsicum annuum seeds (acetone treated) 96.35 Akar et al. (2011)  
Waste biomass of Aspergillus fumigatus 60.6 Wang et al. (2016)  
A. niger powder 29.96 Xiong et al. (2010)  
AL 20.83 Bouhadjra et al. (2021a)  
ALa 50.76 – 
ALc 102.04 – 
Untreated peanut shell (UPS) 30.30 Abbas & Trari (2023d)  
Activated biochar from oil palm 393.97 Jabar & Odusote (2020)  
Agricultural waste products 8.058 Habeeb & Ghawi (2023)  
Agricultural waste products (HCL) 6.514 – 
Agricultural waste products (KOH) 4.17 – 
Native potato peel powder (PP) 85 Bouhadjra et al. (2021b)  
Potato peel powder activated (PPa) 108 – 
Potato peel powder calcined (PPc) 268 – 
Untreated adsorbent 30.40 This study 

Desorption is an unavoidable process and an intermediate stage toward adsorbent regeneration. The latter is an essential step to estimating the reutilization of any adsorbent for industrial applications, owing to the ecological concerns and to needs for sustainable development. Desorption is an inevitable process and an intermediate step toward regeneration of the adsorbent. It is essential for the reuse of any adsorbent for industrial applications, due to ecological concerns and sustainable development needs. Several methods exist for regeneration such as electrochemical, microbiological, thermal, and chemical processes.

We opted for the chemical method because of its low cost; indeed, the regeneration and reuse of adsorbents for other cycles is economically important. The desorption of CB from UPP was evaluated using three different solvents: H2O, NaOH, and HCl. The highest desorption was achieved in HCl solution with a rate of 65.56% against 45.24% (NaOH) and 15.23% (H2O) Figure 15. This result suggests that UPP was suitable for two cycles (R1 = 50.58%, R2 = 32.26%) because of the physisorption character.
Figure 15

Regeneration of the adsorbent (UPP).

Figure 15

Regeneration of the adsorbent (UPP).

Close modal

This study has shown that the UPP prepared can be employed as an effective adsorbent for the removal of CB from aqueous solution. The Langmuir isotherms model provided a better fit of the equilibrium adsorption data one. They gave a maximum adsorption capacity of 30.30 mg/g at temperature 25 °C. The adsorption of CB ions by UPP follows a pseudo-second-order kinetic model, which relies on the assumption that physisorption may be the rate-limiting step. In physisorption, the CB ions are attached to the adsorbent surface by chemical bond and tend to find sites that maximize their coordination number with the surface. The kinetics and thermodynamic data can be further explored for the design of an adsorber for industrial effluents treatment.

The negative values of ΔG° and negative value of ΔH° indicate that the adsorption of CB onto UPP is spontaneous and exothermic over the studied range of temperatures. The negative value of ΔS° states clearly that the randomness increases at the solid–solution interface during the CB adsorption onto UPP, indicating that some structural exchange may occur among the active sites of the adsorbent and the ions.

This study is a tiny batch that gave rise to encouraging results, and we wish to achieve the adsorption tests in column mode under the conditions applicable to the treatment of industrial effluents. Furthermore, the present investigation showed that UPP is a potentially useful adsorbent for metal, acid, and basic dyes. The elimination of CB in the column mode and the degradation of dye by heterogeneous photocatalysis is the logical next step.

The authors gratefully acknowledge support from the Laboratory of Applied Chemistry and Materials (LabCAM), University of M'hamed Bougara of Boumerdes, Avenue de l'Indépendance Boumerdes, 35000, Algeria.

M.A. contributed to preparation, conceptualization, formal analysis, investigation, writing-original draft and methodology. M.T. and M.A. contributed to validation data curation, supervision, writing-review and editing. All authors have read and agreed to the published version of the paper.

The authors disclosed receipt of the following financial support for the research, authorship, and/or publication of this paper.

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

The authors declare there is no conflict.

Abbas
M.
(
2020
)
Performance of apricot stone (ACAS) to remove dyes from aqueous solutions –Equilibrium, kinetics, isotherms modeling and thermodynamic studies
,
Materials Today: Proceedings
,
31
,
437
443
.
Abbas
M.
(
2021b
)
Modeling of adsorption isotherms of heavy metals onto apricot stone activated carbon: Two-parameter models and equations allowing determination of thermodynamic parameters
,
Materials Today: Proceedings
,
43
,
3359
3364
.
Abbas
M.
(
2022a
)
Adsorption of methyl green (MG) in aqueous solution by titanium dioxide (TiO2): Kinetics and thermodynamic study
,
Nanotechnology for Environmental Engineering
,
7
(
1
),
713
724
.
Abbas
M.
(
2022b
)
Removal of basic fuschine in aqueous solution by adsorption process onto PrunusCeracefura (LPC) Kinetic, Isotherm modeling and thermodynamic study
,
Journal of Engineered Fibers and Fabrics
,
17
(
1–10
),
1
10
.
Abbas
M.
(
2022c
)
Removal of Methylene Blue (MB) pollutant from the textile industry by adsorption onto zeolite kinetic and thermodynamic study
,
Journal of Engineered Fibers and Fabrics
,
17
,
1
11
.
Abbas
M.
,
Harrache
Z.
&
Trari
M.
(
2020
)
Mass-transfer processes in the adsorption of Crystal Violet (CV) by activated carbon derived from pomegranate peels: Kinetics and thermodynamic studies
,
Journal of Engineered Fibers and Fabrics
,
15
(
1
),
1
11
.
Aniagor
C. O.
&
Menkiti
M. C.
(
2018
)
J. Environ. Chem. Eng., 6, 2105
.
Bouhadjra
K.
,
Lemlikchi
W.
,
Ferhati
A.
&
Mignard
S.
(
2021b
)
Enhancing removal efficiency of anionic dye (Cibacron Blue) using waste potato peels powder
,
Scientific Reports
,
11
,
2090
.
Cheung
C. W.
,
Porter
J. F.
&
McKay
G.
(
2000
)
Elovich equation and modified second-order equation for sorption of cadmium ions onto bone char
,
Journal of Chemical Technology and Biotechnology
,
75
(
11
),
963
970
.
Chien
S. H.
&
Clayton
W. R.
(
1980
)
Application of Elovich equation to the kinetics of phosphate release and sorption in soil
,
Soil Science Society of American Journal
,
44
,
265
268
.
Deng
H.
,
Lu
J.
,
Li
G.
,
Zhang
G.
&
Wang
X.
(
2011
)
Chem. Eng. J. 172, 326
.
Freundlich
H. M. F.
(
1906
)
Uber dieadsorption ilosungen
,
Zeitschrift für Physikalische Chemie
,
57
,
385
470
.
Habeeb
S. D.
&
Ghawi
A. H.
(
2023
)
Removal of Cibacron blue 3 GA from industrial waste water using agricultural waste products as adsorbent
,
Journal of University of Babylon for Engineering Science (JUBES)
,
31,N5
,
5417
5429
.
Hammeed
B. H.
,
Ahmad
A. L.
&
Latiff
K. N. A.
(
2007
)
Adsorption of basic dye (methylene blue) onto activated carbon prepared from rattan sawdust
,
Dyes and Pigments
,
75
(
1
),
143
149
.
Ho
Y. S.
&
McKay
G.
(
1998
)
Kinetic models for the sorption of dye from aqueous solution by wood
,
Process Safety and Environmental Protection
,
76
(
2
),
183
191
.
Inyinbor
A. A.
,
Adekola
F. A.
&
Olatunji
G. A.
(
2016
)
Water Resour. Ind. 15, 14
.
Irish
M.
&
Irish
G.
(
2000
)
Agaves, Yuccas, and Related Plants, A Gardener's Guide
, Vol
18
.
Portland
:
Timber Press
.
Jawad
A. H.
,
Abdulhameed
A. S.
,
Surip
S. N.
&
Sabar
S.
(
2022
)
Adsorptive performance of carbon modified chitosan biopolymer for cationic dye removal: Kinetic, isotherm, thermodynamic, and mechanism study Int
,
Journal of Environnement Analytical Chemistry
,
102
,
6189
6203
.
Lagergen
S.
(
1898
)
Zurtheorie der sogenannten adsorption gelosterstoffe
,
Kungliga Svenska Vetenskapsakademiens Handlingar
,
24
(
4
),
1
39
.
Langmuir
I.
(
1918
)
The adsorption of gases on plane surfaces of glass, mica and platinum
,
Journal of the American Chemical Society
,
40
,
1361
1403
.
Liu
D.
,
Han
G.
,
Huang
J.
&
Yumning
Z.
(
2009
)
Composition and structure study of naturel Nelumbo nucifera fiber
,
Carbohydrate Polymers
,
75
(
1
),
39
43
.
Singanan
M.
,
Abebaw
A.
&
Vinodhini
S.
(
2005
)
Removal of lead ions from industrial waste water by using biomaterials: A novel method
,
Bulletin of the Chemical Society of Ethiopia
,
19
(
2
),
289
294
.
Temkin
M.
&
Pyzhev
V.
(
1940
)
Kinetics of ammonia synthesis on promoted iron catalysts
,
Acta Physicochimistry.URSS
,
12
,
327
356
.
Vörösmarty
C. J.
,
McIntyre
P. B.
,
Gessner
M. O.
,
Dudgeon
D.
,
Prusevich
A.
,
Green
P.
&
Davies
P. M.
(
2010
)
Global threats to human water security and river biodiversity
,
Nature
,
467
(
7315
),
555
561
.
Wallace
J. S.
(
2000
)
Increasing agricultural water use efficiency to meet future food production
,
Agriculture, Ecosystems & Environment
,
82
,
105
119
.
Weber
W. J.
&
Morris
J. C.
(
1963
)
Kinetics of adsorption on carbon from solution
,
Journal of the Sanitary Engineering Division
,
89
(
2
),
31
60
.
Xiong
X. J.
,
Meng
X. J.
&
Zheng
T. L.
(
2010
)
Biosorption of CI Direct Blue 199 from aqueous solution by nonviable Aspergillus Niger
,
Journal of Hazardous materials
,
175
(
1–3
),
241
246
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).