Open Access
CC BY 4.0 · Sustainability & Circularity NOW 2025; 02: a27596085
DOI: 10.1055/a-2759-6085
Original Article

Harnessing the Adsorption Capacity of Typha domingensis for the Sustainable Elimination of Noxious Bromophenol Blue Dyes from Aqueous Systems

Authors

  • Abul Bashar

    1   Chemistry, Khulna University, Khulna, Bangladesh (Ringgold ID: RIN247289)
  • Al Shariar Hasan

    1   Chemistry, Khulna University, Khulna, Bangladesh (Ringgold ID: RIN247289)
  • Rezaul Haque

    1   Chemistry, Khulna University, Khulna, Bangladesh (Ringgold ID: RIN247289)
  • Palash Kumar Dhar

    1   Chemistry, Khulna University, Khulna, Bangladesh (Ringgold ID: RIN247289)
  • Sagar Kumar Dutta

    1   Chemistry, Khulna University, Khulna, Bangladesh (Ringgold ID: RIN247289)

The study was partially supported by the Research and Innovation Centre, Khulna University, Khulna-9208, Bangladesh, under Grant/Award Number KU/RICKU/04/2000-41.
 


Graphical Abstract

Abstract

Synthetic dyes in industrial wastewater pose serious environmental challenges due to their toxicity and persistence. This study explores a sustainable and affordable biosorbent made from Typha domingensis leaves powder (TDLP) for removing bromophenol blue (BPB) dye from water. The material was characterized using FTIR, SEM-EDS, and pHpzc analyses to understand its surface chemistry, morphology, and charge behavior. Batch adsorption experiments were carried out to evaluate the influence of pH, contact time, dye concentration, and adsorbent dose. Under optimal conditions (pH 2.0, 10 mg L−1 BPB, 0.1 g TDLP per 50 mL, 60 min, 300 K), TDLP achieved a 92% dye removal efficiency. The adsorption process followed the pseudo-second-order kinetic model and fit best to the Langmuir isotherm, indicating monolayer adsorption with a maximum capacity (qm[?down]?>ₘ) of 15.61 mg g−1. Thermodynamic analysis showed that the process was spontaneous and endothermic. The biosorbent retained over 80% of its efficiency after five reuse cycles, confirming its durability. Owing to its natural abundance, low cost, and biodegradability, T. domingensis provides an eco-friendly and circular solution for dye removal, supporting cleaner water management and sustainable environmental practices.


Significance

In this study, a sustainable and low-cost biosorbent was prepared from Typha domingensis plant (TDP) and effectively utilized for the removal of bromophenol blue dye from aqueous solutions. This work contributes to the UN Sustainable Development Goals, particularly SDG 6 (Clean Water and Sanitation) and SDG 12 (Responsible Consumption and Production).

1

Introduction

Water is essential for virtually all living organisms. Clean water is a fundamental necessity for the survival and growth of all life forms. However, due to increasing urbanization, industrialization, population growth, conflicts, climate change, and energy production, the global supply of clean water has become a critical concern of unprecedented magnitude.[1] Currently, one of the most prevalent causes of death globally is water contamination, which results in 2 million deaths every year, and an estimated 1 billion people struggle to find uncontaminated drinking water. The World Health Organization (WHO) predicts that up to half of the world’s population will have limited access to clean water due to climate change, and a recent United Nations report suggests that the world may experience a 40% water shortage in as little as 15 years. As a result, the problem is expected only to worsen.[2] As per the World Health Organization (WHO), 3,900 children die each day due to spread of diseases by contaminated water or poor hygiene.[3] Water contamination is an increasingly serious environmental concern in modern society, resulting in imbalances in the environment and health risks. Given that all living things require water to survive and grow, water contamination is an extremely serious problem. Water contamination arises from a variety of sources, including radioactive waste disposal, eutrophication, atmospheric deposition, oil spills, sewage discharge, industrial waste dumping, global warming, marine dumping, and more.[4] [5] Numerous chemicals are being utilized or released, resulting in increasingly severe pollution. The quality and taste of water can be diminished by pesticides, pharmaceuticals, heavy metals, oils, detergents, industrial chemicals, and dyes. Dye pollutants in freshwater can originate from various industries, including textiles, pharmaceuticals, food, leather, paint, and varnish.[6]

Dyes are among the most harmful chemical compounds to the environment. They are widely employed across multiple industries and are regarded as major sources of water contamination. There are almost 100,000 different types of dyes available globally, and over 10,000 tons are produced annually. Approximately 15% of all dyes are released into wastewater during dyeing and production processes, mostly because the dyes are not completely fixed onto fabrics.[7] The discharge of highly colored wastewater from dyeing industries not only poses hazards to aquatic life but also reduces the aesthetic value of affected water bodies. If dyes degrade in sediment, incomplete bacterial breakdown can produce toxic amines, which are potentially dangerous.[8] Water contamination by dyes is a worldwide issue due to their persistence, toxicity, and health impacts. Many pigments and dyes are harmful to both aquatic life and humans, causing cancer, mutagenic effects, and teratogenesis.[9] Water-soluble dyes are hazardous to microorganisms and inhibit photosynthetic activity and the growth of various aquatic organisms.[10]

Dyes are categorized based on their applications and chemical structures.[11] Azo dyes represent a large group of synthetic coloring agents containing azo functional groups (–N=N–) and are classified as anionic, cationic, or nonionic. Azo dyes include aromatic rings and –N=N– groups, making them highly hazardous, carcinogenic, and teratogenic. They damage ecosystems and harm living organisms.[12] Among the industries with the highest chemical intensity, the textile industry is a major contributor to water pollution. Large amounts of dyes are released as wastewater during various stages of textile processing. One commonly used compound in this sector is sulfonated azo dye, an aromatic sulfonate. Because of their high mobility in aquatic systems, these dyes can easily evade water treatment processes and contaminate surface waters.[13] Azo dyes are both acutely and chronically hazardous. Contact with them can cause skin and eye irritation, while prolonged exposure may lead to toxicity, carcinogenicity, and mutagenicity in humans.[14] Up to 20% of the massive quantities of azo dyes used in the textile sector are eventually released into the environment as wastewater, causing high levels of cytotoxicity and mutagenic effects on human health and aquatic species.[15] Due to their toxicity and visual impact, dye-containing effluents have recently gained significant attention. To mitigate their environmental effects, various treatment techniques are employed, including ozonation, microbial degradation, adsorption, reverse osmosis, ion exchange, coagulation/flocculation, photocatalytic decolorization, wet air oxidation, sono-chemical, and electrochemical methods.[16] [17] [18] Researchers have explored advanced materials to address such pollutants. Metal-organic frameworks (MOFs) and their derivatives have shown promise due to their high porosity, tunable functionalities, and strong adsorption capacities for emerging contaminants. MOF-derived carbon materials,[19] though efficient in water purification, require high-temperature pyrolysis and complex post-treatments for structural optimization. Another approach involves bio-electrochemical systems such as microbial fuel cells (MFCs), where bio-cathode materials have been investigated for simultaneous energy generation and wastewater treatment. Although innovative, these systems are better suited to continuous-flow conditions and demand significant infrastructure and maintenance, making them less practical in low-resource settings.[16] [17] [18] Nanocomposites synthesized from waste materials, such as chitosan–magnesium oxide derived from shrimp shells and MgO nanoparticles supported on carbon,[16] [17] [20] have demonstrated high adsorption capacities and effective dye removal. However, these methods often involve multistep synthesis, chemical treatments, and nonrenewable precursors. The high initial and operating costs further limit their large-scale feasibility. Conversely, ion exchange and reverse osmosis allow for pollutant recovery along with removal, but their high capital and operational costs make them economically less feasible.[21] Among the numerous dye-removal techniques, adsorption has been identified as the most efficient and practical method.[22]

Adsorption is a mass transfer process involving the adhesion of gases or solutes onto solid or liquid surfaces. It occurs when molecules or atoms on a solid surface retain residual surface energy due to imbalanced forces. These forces attract certain substances that collide with the surface, causing them to adhere. Recently, the scientific community has expressed strong interest in adsorption-based wastewater treatment.[23] [24] The adsorption technique is considered promising for dye removal due to its practicality, simplicity, and ease of operation.[25] Many researchers have investigated the efficiency of various biological wastewater management processes to eliminate high-risk pollutants (HRPs) and minimize environmental hazards. The material that accumulates on the solid surface during adsorption is called the adsorbate, while the solid itself is referred to as the adsorbent.[26] Adsorption thus describes the surface accumulation of an adsorbate on an adsorbent. Adsorbents are categorized as either natural or synthetic. Charcoal, clays, clay minerals, zeolites, and ores are examples of natural adsorbents.[27] Natural materials are generally inexpensive and abundant, with considerable potential for surface modification to improve adsorption performance. Synthetic adsorbents include materials derived from agricultural and household waste, industrial residues, sewage sludge, carbon nanomaterials, and synthetic polymers.[28] [29] [30]

Among the most widely used adsorbents for pollutant removal is activated carbon.[31] Owing to its exceptional adsorption capacity, reactive surface, porosity, inertness, and thermal stability, activated carbon is extensively employed. It exists in two forms: powdered activated carbon (PAC) and granular activated carbon (GAC).[32] Although GAC is more expensive than PAC, it is easier to regenerate.[33] Because of its versatility, researchers have attempted to reduce costs by producing activated carbon from low-cost sources or through surface modification.[34]

Various biosorbents have also been examined for toxic dye removal. Naturally occurring plant biomass offers a cost-effective alternative for removing dyes from synthetic and industrial effluents.[35] In the present study, the biosorption capacity of T. domingensis leaf powder was evaluated for removing harmful dyes from water.[36] T. domingensis was selected because it is abundant and resilient in polluted aquatic environments, often growing in industrial effluent lakes.[37] The effects of several parameters on dye biosorption using T. domingensis leaf powder (TDLP) were explored.[38] [39] Commonly known as Southern cattail (locally “Hogla”), T. domingensis (family: Typhaceae) is a tall aquatic grass that can grow up to three meters high in clusters. The perennial plant is mainly found in southern Bangladesh and the Sundarbans but also occurs along rivers, canals, and other water bodies, including around Khulna University.

Bromophenol blue (BPB) is a toxic, non-biodegradable, and mutagenic azo dye that threatens both ecosystems and human health. It is a widely recognized anionic dye used in several industries, including textiles, paper, printing, pharmaceuticals, food, and research laboratories. Other sectors such as mining, tanneries, metal smelting, and battery manufacturing also contribute to environmental contamination with BPB. Because certain components of this dye resist biological degradation, its presence in water poses risks to all living organisms, including microorganisms and humans.[40] [41]

This study investigates the potential of T. domingensis leaf powder as a biosorbent for removing BPB from aqueous media. The selection of T. domingensis is based on its natural abundance, pollution resistance, and reported dye biosorption efficiency. Key parameters such as pH, contact time, dye concentration, and adsorbent dosage were evaluated to understand its biosorption behavior. This work aims to develop a low-cost, renewable, and effective biosorbent for sustainable wastewater treatment.


2

Materials and Methods

2.1

Materials

The plant sources that were used in this study include T. domingensis, commonly known as Southern cattail and the leaves of the plant were collected from Khulna University’s campus, Bangladesh. The leaves were transformed into activated charcoal in the next step. Activation of the carbonized samples was achieved through the use of lemon fruits. BPB, an anionic dye frequently used in textile industries, was chosen as the model pollutant for adsorption studies. Analytical grade BPB was purchased from Sigma-Aldrich, which was used in its original form without further treatment. The chemicals and materials used were of analytical quality and did not require any further purification before use. To perform analytical assessments, sodium hydroxide (NaOH) and hydrochloric acid (HCl), both of analytical quality, were used to adjust pH levels. De-ionized water was utilized consistently across all experiments for the preparation of every solution.


2.2

Preparation of Adsorbent

The leaves of T. domingensis were collected, ensuring they were mature and healthy. To get rid of any potential sand or dirt, the fruit shells were repeatedly cleaned with both tap water and de-ionized water. After being cleaned, the shells were dried in the sun for 8 h. The leaves were boiled for 30 min to extract the coloring components (tannic, pigment, and chlorophyll), after which they were allowed to dry in the air and then baked at 110 °C for an entire night. The leaves were then manually crushed with a mortar and pestle until they were powdered ([Fig. 1]). After filtration, the powder was left to dry at 110 °C for 24 h. After removing the bigger particles with a 100 μm filter, the powder was utilized as a biosorbent without undergoing any chemical processing.[1]

Zoom
Fig. 1 Preparation of dry TDLP.

2.3

Preparation of Bio-activators

The dried sample was then placed in porcelain crucibles as well as carbonized for 2 h at 650 °C in a Muffle furnace. With porcelain crucibles as holders, 300 g of feedstock per batch was carbonized in the Muffle furnace for 2 h at 650 °C. The two feedstock materials provided a combined total of 500 g of carbonized materials for each case. To achieve 500 g of carbonized products for each feedstock, the carbonization process was carried out repeatedly at 500 °C.[42]


2.4

Activation of Carbonized Compounds Using Lemon Juice Extract

The juice of lemon fruit was taken out, and the pH was determined. Glass containers were used to store the collected juice, which was acidic. The carbonized samples were combined with 250 mL of the extracted lemon juice. After stirring them until a smooth mixture was formed, they were left to stand at room temperature for a full day. The next step involved filtering and drying the sample in an oven set at 110 °C for a 24-hour period. After being heated for an hour at 450 °C in a Muffle furnace, the samples were cooled and repeatedly cleaned until they reached a neutral state. The last sample was placed in an oven at 110 °C and dried until its weight became steady.[42]


2.5

Characterization

In this investigation, multiple characterization techniques were applied comprehensively to analyze the adsorbent properties as well as surface morphology of T. domingensis. The pH at which the adsorbent has a neutral surface charge, or the point of zero charge (pHpzc), was assessed to explore the adsorbent’s surface charge behavior as a function of the solution’s pH.[43] The process involved introducing measured amounts of T. domingensis to solutions at various initial pH levels and then tracking the final pH to find the point where the surface of adsorbent became neutrally charged. Fourier Transform Infrared Spectroscopy (FTIR) was applied to examine the functional groups present on the surface of T. domingensis, providing insight into potential adsorption sites and mechanisms by examining the shifts in characteristic peaks before and after dye adsorption. Combining Scanning Electron Microscopy with Energy Dispersive X-ray Spectroscopy (SEM-EDS) provides detailed images of the morphology of the surface and its elemental composition. The images of SEM reveal the surface texture and porosity, while EDS analysis confirms the presence of key elements, indicating successful adsorption of BPB onto the adsorbent. These combined techniques ensured a thorough characterization of T. domingensis, enabling a deeper understanding of its adsorption capacity and mechanism for the elimination of BPB dyes from aqueous media.


2.6

Preparation of the Dye Solution

To prepare a solution of BPB dye with a concentration of 500 mg/L, 0.125 g of the dye was mixed in water, which was deionized, and adjusted the volume to 250 mL. The solution was contained in a volumetric flask to enable the preparation of more diluted solutions. Various dye concentrations were generated by diluting the stock solution. Standard preparation involved solutions with concentrations of 10, 20, and 30 mg/L. These solutions were used to study different parameters of adsorption.


2.7

Batch Adsorption Studies

After preparing the stock solution of dyes, 50 mL of dye solutions were taken in beakers. To regulate the dye solution’s pH, 0.1 M hydrochloric acid and 0.1 M sodium hydroxide solutions were also taken. When the amount of the adsorbent taken was 0.1 g and the dye mixture solution concentration was 500 ppm, the pH effect was ascertained. Initially, 0.1 M NaOH and 0.1 M HCl were utilized to adjust the pH (2, 4, 6, 8, 10, and 12) solutions. While the adsorbent dose was at 0.1 g, varied dye concentrations were taken from 10–30 ppm, with 10 ppm intervals, to investigate the impact of the dye solution concentration on adsorbents. Similarly, to investigate the impact of the contact duration of the dye solutions on the adsorbents, the adsorbent dose as well as varied dye concentrations were kept at 0.1 g and 10, 20, and 30 ppm, respectively, and the subsequent contact time was 10–60 min at 10-min intervals. The TDLP charcoal dose effect on the dye solution was investigated by limiting the contact time and concentration to 60 min and 10 ppm, respectively. This procedure was repeated for 0.01, 0.03, 0.05, 0.08, 0.1, and 1.2 g with the same parameters.

The dye removal percentage (%) was estimated using the following relationship:

Removal efficiency % = C i C f C i × 100 %

where the initial dye concentrations are denoted as Ci in mg/L and the final dye concentrations as Cf in mg/L. The dye uptake per gram of adsorbent, in milligrams per gram, was determined through the relationship specified below.

q e = C 0 C e V m

where the initial and equilibrium dye concentrations are Co and Ce in mg/L, respectively, the volume of the solution in liters is indicated by V, and the mass of the adsorbent in grams is represented by m.



3

Results and Discussion

3.1

Fourier Transforms Infrared Spectroscopy (FT-IR) Analysis

The T. domingensis biosorbent was characterized using FT-IR analysis both before and after the adsorption of the BPB dye as shown in [Fig. 2]. This examination revealed important details about the relationship between the dye and T. domingensis functioning as the adsorbent. Initially, the FT-IR spectra of T. domingensis revealed characteristic peaks corresponding to various types of functional groups, including hydroxyl (–OH), carbonyl (C=O), and carboxyl (–COOH) groups. Specifically, a broad peak at approximately 3200 cm−1 indicated bonded (O–H) of carboxylic acid, while peaks at around 2360, 1740, and 1086 cm−1 were attributed to O=C=O stretching of carbon dioxide, C=O stretching, and –COO– asymmetric stretching vibrations, respectively.

Zoom
Fig. 2 Combination FTIR spectrum of TDLP before and after adsorption.

Following the adsorption of BPB, noticeable shifts and changes in peak intensities were observed in these spectra, indicating these functional group involvement in the process of adsorption. The shift of the –OH stretching peak to a lower wavenumber signals hydrogen bond interactions between the T. domingensis hydroxyl groups and the dye molecules. Additionally, a slight shift in the stretching peak of C=O implied the role of carbonyl groups in the mechanism of adsorption.

These spectral modifications support the hypothesis that multiple functional groups on T. domingensis’’ surface bind actively to the BPB dye. The appearance of new peaks or intensified existing peaks in the FT-IR spectra postadsorption further suggests the occurrence of new interactions or chemical bonds between the dye molecules and the adsorbent. These findings align with previous research emphasizing the significance of hydroxyl, carbonyl, and carboxyl groups in the adsorption of organic dyes onto plant-based adsorbents.[37] [44]


3.2

The Zero Point Charge

The pHZPC, or point of zero charge, is the pH at which the surface of a substance has a neutral charge, meaning it is electrically neutral. This happens when the amount of positively charged sites matches the negatively charged ones, leading to a net zero charge density. The pHZPC is most important in adsorption processes. If the pH is less than the pHZPC, the surface of the adsorbent takes on a positive charge, while a pH greater than the pHZPC results in a negative charge surface.[45]

The determination of TDLP’s pHZPC was done through the solid addition method. In this method, 50 mL of 0.1 M NaCl solution at various pH values (ranging from 2 to 12) was placed into beakers, and 0.1 g of TDLP adsorbent was subsequently added. Adjustments to the pH of the solutions were made using 0.1 M HCl or NaOH. At room temperature, the solutions were mixed continuously for 48 h, after which the final pH was recorded. The initial pH (pH i ) and the corresponding change in pH (ΔpH) were represented in a graph ([Fig. 3] ). The intersection point of this graph indicates the pHZPC of the TDLP surface, which was found to be 7.54. This means that for pH levels below 7.54, the TDLP surface is positively charged, and for pH levels above 7.54, it is negatively charged. Since BPB is an acidic dye (also known as an anionic dye), it is adsorbed onto the positively charged surface of TDLP.[46]

Zoom
Fig. 3 Determining the point of zero charge (pHzpc) of TDLP.

3.3

SEM–EDS Analysis

The SEM-EDS analysis offers essential information about the structural characteristics and compositional changes of the adsorbent material. The SEM images reveal that the surface morphology of T. domingensis is notably porous and fibrous, which is beneficial for adsorption processes because of the increased area of surface available for dye interaction. After the process of adsorption, the SEM images show ([Fig. 4]) a significant alteration in surface texture, indicating the adherence of dye molecules to the adsorbent surface.[47]

Zoom
Fig. 4 SEM images of (a) TDLP before adsorption and (b) TDLP after adsorption.

The analysis of Energy Dispersive X-ray Spectroscopy (EDS) further corroborates these findings by confirming the presence of elemental signals characteristic of BPB on the adsorbent surface after adsorption. Prior to adsorption, the EDS spectrum ([Fig. 5]) predominantly displays peaks corresponding to the inherent elements of T. domingensis, such as carbon and oxygen. After adsorption, new peaks corresponding to sulfur and bromine, which are components of BPB, are observed, confirming the successful attachment of the dye to the adsorbent material. This combined SEM-EDS analysis not only validates the physical adsorption of BPB on to T. domingensis but also highlights the adsorbent’s potential effectiveness and structural adaptability in removing toxic dyes from aqueous solutions.[1]

Zoom
Fig. 5 EDS images of (a) TDLP before adsorption and (b) TDLP after adsorption.

3.4

Batch Adsorption Studies

3.4.1

Effect of pH

In the batch studies for the research, the effect of pH on the adsorption efficiency is systematically investigated. The dye adsorption capacity is tested at varying pH levels to identify the best conditions for effective removal. The results demonstrate that the adsorption capacity fluctuates notably with pH changes, indicating the impact of pH on both the adsorbent’s surface charge and the ionization behavior of BPB.

When the pH is lowered, the surface of T. domingensis gains a positive charge, which strengthens the electrostatic attraction to the negatively charged dye molecules, resulting in increased removal efficiency. As the pH increases, the surface charge of the adsorbent changes, leading to reduced electrostatic attraction and a lower adsorption efficiency. Dye removal reaches its peak at a specific acidic pH, emphasizing the fundamental role pH plays in the adsorption process.[48] [49]

In other words, lower pH levels tend to reduce the effectiveness of cationic dye removal but improve the removal efficiency for anionic dyes. Conversely, higher pH levels tend to reduce the percentage anionic dye removal but improve the removal efficiency for cationic dyes.[50] The percentage of removal was reduced from 84.67% to 4.79 % with the rise of pH from 2 to 12. At pH 2, the highest removal was observed ([Fig. 6]). A removal efficiency of 84.67% was achieved at a pH value of 2. This study confirms that pH is a critical element in refining adsorption processes for the successful extraction of harmful dyes from water. Understanding the pH dependence of the adsorption process allows for the fine-tuning of operational parameters, enhancing the practical applicability of T. domingensis as a sustainable and efficient adsorbent for wastewater treatment.

Zoom
Fig. 6 pH Effect of BPB dye removal.

3.4.2

Effect of Adsorbent Dose

Various doses of T. domingensis adsorbent were evaluated in batch studies to assess their effect on the removal efficiency of BPB dyes and to identify the optimal amount needed for peak performance. The studies were conducted by adding different weights of T. domingensis to a fixed volume of the dye solution and agitating the mixture for a predetermined time. To evaluate the results, the dye concentration remaining in the solution after adsorption was measured. The findings are displayed in [Fig. 7]. The trials were carried out by adding varying amounts of adsorbent 0.01, 0.03, 0.05, 0.08, 0.1, and 1.2 g in 50 mL of dye solution. Temperature, pH, and contact time were the other variables that remained constant. As the adsorbent dose was increased from 0.01, 0.03, 0.05, 0.08, 0.1, to 1.2 g, the removal percentage increased from 81.52%, 86.41%, 87.50%, 91.30%, 91.85%, to 91.85%, respectively. Greater dye removal was attained with greater adsorbent doses, as was predicted. The greater availability of active sites for the same number of adsorbate molecules may be the reason for the higher clearance at higher doses.[51]

Zoom
Fig. 7 Adsorbent dosage effect on BPB dye removal.

The percentage removal of BPB was calculated for each adsorbent dose, revealing a clear relationship between the adsorbent dose and the removal efficiency. As the dose of T. domingensis increased, the percentage of dye removal also increased, indicating enhanced adsorption capacity. This trend continued until a certain point where further increases in the adsorbent dose resulted in only marginal improvements in dye removal efficiency, suggesting the saturation of available adsorption sites on the T. domingensis surface.[52]

Broadly, these findings highlight the importance of optimizing the adsorbent dose to achieve maximum removal efficiency while minimizing the amount of adsorbent used, thus making the process more cost effective and sustainable. The study underscores the potential of T. domingensis as a highly effective natural adsorbent for the treatment of dye-contaminated wastewater, contributing to environmental sustainability and pollution.


3.4.3

Effect of Concentration and Contact Time

The adsorption efficiency of T. domingensis for BPB was systematically analyzed by modifying the initial dye concentration while maintaining constant experimental conditions. After 60 min, an increase in dye concentration was observed from 10 to 30 mg/L, resulting in a decrease in adsorption efficiency from 91.85% to 43.24% ([Fig. 8]). After reaching equilibrium, the dye removal efficiency did not significantly alter. As the adsorbent dose was consistent, the dye molecules were able to attach to enough active sites at low concentrations, improving the efficiency of dye removal. Additionally, there were a set number of binding sites in the adsorbent, and these sites were satiable at a specific concentration. Therefore, lower dye removal efficiency resulted from more dye molecules remaining unabsorbed due to surface saturation as the concentration was increased.[53]

Zoom
Fig. 8 Contact time and initial dye concentration effect on BPB dye removal.

The findings revealed that the starting concentration of BPB played a crucial role in determining the adsorption rate. At lower concentrations, the adsorption sites on T. domingensis were more readily accessible, facilitating rapid uptake of dye molecules during the early stages of contact time. However, as the concentration increased, these adsorption sites became progressively saturated, leading to a corresponding decline in the adsorption rate.[54]

Further analysis of the relationship between concentration and contact time was conducted to ascertain the equilibrium time necessary for achieving maximum adsorption. It was noted that as concentration increased, more time was needed to reach equilibrium, likely due to greater competition among dye molecules for the limited adsorption sites.[55]

This study underscores the critical importance of optimizing both the dye concentration and the contact time to maximize adsorption performance. A comprehensive understanding of these factors is crucial for the practical application of T. domingensis for treating wastewater that holds different concentrations of hazardous dyes.



3.5

Adsorption Kinetics

Adsorption kinetics were studied to investigate the mechanism involved in BPB dye molecule adsorption. This analysis focused on the adsorption process to evaluate the rate at which adsorption takes place. The study of kinetics was modeled using both pseudo-first-order and pseudo-second-order equations. The kinetics study of BPB dye adsorption process by TDLP was performed at dose 0.1 g/50 mL, 300 K temperature, pH of 2, 60 min contact time with various concentrations (10, 20, and 30 mg/L).

3.5.1

Pseudo-First-Order Kinetic Model

The adsorption process of BPB dye on T. domingensis was analyzed through the pseudo-first-order model.

The pseudo-first-order equation, as outlined by Lagergren in 1898, is represented as

d q t q t = K 1 q e q t

where, the capacity of adsorption (mg/g) is represented as qe , the dye adsorption at time t (mg/g) represented as qt , and the adsorption rate constant at equilibrium represented as K 1.[56] [Table 1] shows the calculated K 1, qe , qt , and R 2 values obtained from these parameters.

Table 1

Kinetic parameter of BPB adsorption on to TDLP for pseudo-first-order kinetics

Initial concentration (mg/L)

Q e (exp) (mg/g)

k 1 (min−1)

R 2

10

2.969

−0.0011

0.78023

20

6.011

−0.0009

0.99339

30

3.216

−0.0003

0.78183

Under the boundary condition, the equation can be transformed linearly using the following expression.

ln q e q t = ln q e K 1 2.303 t

By plotting the values of log (qe qt ) against t, a straight line is formed, from which k 1 and qe can be determined using the slope and intercept. The results of log (qe qt ) at varying concentrations of 10, 20, and 30 mg/L are given in [Fig. 9] for adsorbent TDLP. The correlation coefficient was about 0.78023, which is much lower than 0.99, indicating that the pseudo-first-order model is not applicable for this study.[57] [58]

Zoom
Fig. 9 Pseudo-first-order kinetics at different concentrations for BPB.

3.5.2

Pseudo-Second-Order Kinetic Model

The adsorption behavior of BPB dye on T. domingensis is effectively characterized by the pseudo-second-order kinetic model. In this framework, it is assumed that the rate-limiting step could involve either chemical adsorption or chemisorption, with the adsorption capability proportional to the number of active sites on the adsorbent.

The pseudo-second-order equation is expressed as

d q t d t = K 2 q e q t 2

where, the capacity of adsorption (mg/g) is represented as qe , the dye adsorption at time t (mg/g) represented as qt , and the pseudo-second-order kinetics equilibrium rate constant (g/mg/min) represented as K 2.[59] [Table 2] shows the calculated K 2, qe , qt , and R 2 values obtained from these parameters.

Table 2

Kinetic parameter of BPB adsorption on to TDLP for pseudo-second-order kinetics

Initial concentration (mg/L)

Q e (exp) (mg/g)

k 2 (g mg min−1)

R 2

10

3.829

0.02626

0.96605

20

8.443

0.00352

0.98722

30

3.756

0.0185

0.96999

Under the boundary condition, the equation can be transformed linearly using the following expression.

t q t = 1 K 2 q e 2 + t q e

By graphing t/qt in relation to t, a correlation diagram is obtained, from which K 2 and qe values can be calculated using the slope and intercept. The relationship between t/qt and t yields a linear line, as shown in [Fig. 10].

Zoom
Fig. 10 Pseudo-second-order kinetics at different concentrations for BPB.

The equilibrium adsorption capacity, denoted as q e, was assessed at concentrations of 10, 20, and 30 mg/L. The values obtained from the experiments and calculations for the equilibrium adsorption capacity, qe were nearly the same for TDLP adsorbents. The correlation coefficient R 2 for TDLP approached 0.99, signifying its capability to explain BPB adsorption accurately.[60]

The pseudo-second-order model tends to show a more elevated correlation coefficient (R 2) relative to the pseudo-first-order model, pointing to the fact that BPB adsorption on T. domingensis is mainly regulated by chemisorption, which involves the exchange or sharing of electrons between the adsorbent and adsorbate. The investigation into T. domingensis’s capacity to adsorb BPB from aqueous solutions reveals that the pseudo-second order kinetic model offers significant insights. The agreement between the model and the experimental findings indicates that chemical bonding plays a more significant role in the adsorption process than simple physical interactions. This evidence is necessary for analyzing the effectiveness and sustainability of T. domingensis as an eco-friendly adsorbent, highlighting its potential to form stable chemical bonds with harmful dyes, thereby effectively removing them from water.[61] [62]



3.6

Isotherm of Adsorption

The process of adsorption can be effectively described through the important role of the adsorption isotherm. It is important to have a clear understanding of adsorption. Analyzing the mechanism by which adsorbate molecules attach to the adsorbent is essential for understanding the process. The adsorption isotherm was evaluated through the application of the Langmuir and Freundlich isotherm models. The assumption of monolayer adsorption in the Langmuir isotherm suggests that active sites are homogeneously distributed across the adsorbent’s surface. The assumption of multilayer adsorption in the Freundlich isotherm suggests that active sites are heterogeneously distributed across the adsorbent’s surface. This suggests that the surface can support multiple layers of adsorbate molecules, resulting in a more intricate and diverse adsorption process.[63] [64]

3.6.1

Langmuir Isotherm

As outlined by the Langmuir isotherm, maximum adsorption is observed on a monolayer surface that is saturated, containing a fixed number of equivalent sites and consistent adsorption energy. Adsorbents do not move across the plane of the surface.[65]

The linearized expression of the Langmuir equation is shown as,

1 q e = 1 q m + 1 K L q m C e

where, the equilibrium adsorption capacity (mg/g) is represented as qe , the maximum adsorption capacity of the adsorbent (mg/g) represented as qm , the Langmuir adsorption constant (L/mg) represented as KL , and the equilibrium adsorption concentration (mg/L) represented as Ce .[66] A linear straight line is observed when plotting 1/qe versus 1/Ce , as shown in [Fig. 11], from which KL is derived via the slope and intercept.

Zoom
Fig. 11 Langmuir adsorption isotherm for the TDLP.

The fundamental attributes of the Langmuir isotherm model are represented by the separation factor (RL ), which is a dimensionless parameter expressed by the following equation.

R L = 1 1 + K L C o

where Langmuir constant (L/mg) is KL and the initial MB concentration (mg/L) is C o.[67] When RL is more than 1, the adsorption process is not favorable. RL values between 0 and 1 indicate favorable adsorption. RL equal to 1 shows linear adsorption, and RL equal to 0 points shows irreversible adsorption.

Based on [Table 3], the correlation coefficient R 2 of 0.9819 for TDLP strongly suggests that the Langmuir adsorption isotherm is favorable and effectively describes the adsorption process. The RL values, lying between 0 and 1, support the favorable conditions of the adsorption process.

Table 3

Parameters of Langmuir Adsorption Isotherm for BPB on to TDLP at 300 K

Temperature (K)

Langmuir parameters

q m

K L

R L

R 2

300

15.61

0.29282

0.3415

0.9819

313

6.05

0.51098

0.0978

0.9994

323

1.08

3.20488

0.0104

0.9876


3.6.2

Freundlich Isotherm

The Freundlich isotherm explains the adsorption process on a heterogeneous surface where the heat of adsorption is not uniformly distributed. The Freundlich isotherm can be described using the following equation.[68]

ln q e = 1 n ln C e + ln K F

where the equilibrium adsorption capacity (mg/g) is represented as qe ; the equilibrium adsorption concentration (mg/L) represented as Ce ; the Freundlich isotherm constants mg/g (L/mg) represented as KF ; and the adsorption intensity represented as n. When 1/n lies between 0 and 1, adsorption is spontaneous; at 1/n = 1, it becomes irreversible; and if 1/n exceeds 1, the process becomes difficult to carry out.

The plot of ln qe against ln Ce in [Fig. 12], based on the experimental data from [Table 4], indicates that the n value for TDLP is 3.934 from the Freundlich isotherm. This suggests that the adsorption process is spontaneous and shows good characteristics.

Zoom
Fig. 12 Freundlich adsorption isotherm for the TDLP.
Table 4

Parameters of Freundlich Adsorption Isotherm for BPB on to TDLP at 300 K

Temperature (K)

Freundlich parameters

K F

n

R 2

300

2.673

3.934

0.9358

313

1.773

4.701

0.983

323

1.42

2.851

0.9942

[Tables 3] and [4] show that the R 2 value is more aligned with the Langmuir isotherm, making it preferable over the Freundlich isotherm. The highest adsorption capacity recorded for TDLP was 15.61 mg/g, demonstrating that the experimental adsorbent possesses outstanding adsorption characteristics. In [Table 5], the performance of BPB dye adsorption on various surfaces of adsorbents is illustrated. Following the summarization of the adsorption data, it was determined that the BPB adsorption on to TDLP was consistent with a monolayer model.

Table 5

Review of the maximum adsorption capabilities of a range of adsorbents for BPB dye

Adsorbent

Adsorption isotherm

q e (mg/g)

Reference

Chitin nanoparticles

Langmuir

22.72

[69]

Flamboyant pods (FBP)

Langmuir

7.58

[70]

Polymer clay

Freundlich

10.78

[71]

Astragalus bisulcatus

Langmuir

23.45

[72]

Sorel’s cement

Freundlich

4.88

[73]

Musa acuminata peel

Freundlich

6.04

[74]

Polymeric gel

Langmuir

2.98

[75]

Solanum tuberosum peel

Freundlich

8.157

[76]

Activated charcoal

Langmuir

0.081

[77]

TDLP

Langmuir

15.61

This work



3.7

Thermodynamic Studies

Thermodynamic studies are pivotal for understanding the attributes and effectiveness of the adsorption method, particularly when it comes to eliminating BPB dyes using T. domingensis. The van’t Hoff equation is useful for determining thermodynamic parameters, including the Gibbs free energy change (ΔG°), enthalpy change (ΔH°), and entropy change (ΔS°). The values of these thermodynamic parameters are contingent upon temperature and can be calculated using the subsequent equation,[78]

ln K d = S ° R H ° RT

where the thermodynamic equilibrium constant is represented as Kd , the universal gas constant (8.314 J/mol/K) represented as R, the solution temperature (K) represented as T, the entropy change (kJ/mol/K) represented as ∆S°, and the enthalpy change (kJ/mol) represented as ∆H°. The values of Gibbs free energy (∆G°), entropy (∆S°), and enthalpy (∆H°) are presented in [Table 6].

Table 6

Parameters of thermodynamic for adsorption of BPB on to TDLP

Initial concentration (mg/L)

H° (kJ/mol)

S° (kJ/mol)

ΔG° (300 K)

ΔG° (313 K)

ΔG° (323 K)

10

58.42

103.453

−0.237

−1.380

−2.667

Plotting ln Kd versus 1/T results in a straight line as illustrated in [Fig. 13]. Calculations for ∆H° and ∆S° can be made using the slope and intercept. The exothermic nature of the adsorption process is indicated by a negative ∆H°, while a positive ∆H° points to an endothermic process. The change in Gibbs free energy (ΔG°) reflects whether the adsorption process occurs spontaneously or not. A negative value of ΔG° at various temperatures suggests that the adsorption of BPB on to T. domingensis is spontaneous. The standard Gibbs free energy change, ΔG°, is usually determined through the following equation[79]:

Zoom
Fig. 13 Plot of ln Kd vs. 1/T for evaluation of thermodynamics parameters for the adsorption of BPB onto TDLP.
G ° = RT ln K d

From the experimental data given in [Table 6], ∆G° values decrease with increasing temperature and is a negative value. This illustrates that the adsorption process is characterized by spontaneity. Here, the determined ∆H° value of TDLP is 58.42. The positive value of ∆H° indicates that the adsorption process is inherently endothermic. ΔS° was found to be positive, indicating an increase in randomness at the solid–liquid interface during the dye adsorption process. This may be attributed to the desolvation of dye molecules and the formation of more disordered surface interactions. Additionally, a positive ∆S° suggests that there is an increase in randomness as temperature rises.



4

Reusability

The reusability of an adsorbent is the key indicator of its practical feasibility for wastewater treatment applications. [Fig. 14] illustrates the performance of T. domingensis -derived biosorbent (TDLP) over five consecutive adsorption–desorption cycles. The initial removal efficiency of 92% in the first cycle gradually decreased to 90%, 87%, 84%, and finally to 80% in the fifth cycle. This moderate decline in efficiency may be attributed to factors such as partial saturation of the active binding sites, incomplete desorption of dye molecules during regeneration, or slight degradation of the biosorbent’s surface functionality over repeated usage. Notably, the material retained over 80% removal efficiency after five cycles. The time interval between each cycle was approximately 24 h, allowing for desorption, washing, and complete drying of the biosorbent. Compared to synthetic materials, which may offer higher stability but at greater cost and environmental burden, the biosorbent derived from T. domingensis offers a low-cost, biodegradable, and moderately durable alternative for dye removal. Future studies may focus on improving the regeneration protocol or chemically modifying the biosorbent to enhance its long-term reusability.

Zoom
Fig. 14 Reusability Study for the TDLP.

5

Conclusion

This study demonstrates that TDLP is an effective, low-cost, and eco-friendly material for removing BPB dye from water. The adsorption process showed 92% efficiency under optimal conditions and follows the pseudo-second-order kinetic and Langmuir isotherm models, indicating a monolayer chemisorption mechanism. Thermodynamic results confirmed that the process was spontaneous and endothermic, while reuse experiments showed that TDLP maintained over 80% of its efficiency after five cycles. These findings highlight the potential of T. domingensis as a renewable biosorbent for wastewater treatment. Converting this abundant aquatic plant into a functional adsorbent supports circular economy principles by turning natural biomass waste into a valuable material for sustainable water purification and environmental protection.



Contributors’ Statement

M.A.B.: Formal analysis, Software, Investigation, Validation, Writing-original Draft. A.S.H.: Writing-Review & Editing. M.R.H.: Conceptualization, Study design, Fund acquisition, and project administration. P.K.D.: Visualization, Software, Writing-Review & Editing. S.K.D.: Conceptualization, Methodology, Formal analysis, Writing-Review & Editing, Visualization, Software, Supervision.

Conflict of Interest

The authors declare that they have no conflict of interest.

Acknowledgement

The authors express their sincere gratitude to the Chemistry Discipline, Khulna University, Khulna-9208, for providing the necessary laboratory facilities.

  • References

  • 1 Dutta SK, Amin MK, Ahmed J, Elias M, Mahiuddin M. S Afr J Chem Eng 2022; 40: 195-208
  • 2 Sun DT. et al. ACS Cent Sci 2018; 4 (03) 349-356
  • 3 Lee A, Elam JW, Darling SB. Membrane Materials for Water Purification: Design, Development, and Application. Royal Society of Chemistry; 2016.
  • 4 Syed A, Kumar G, Tonu NT, Chakrabarty S, Mahiuddin MD, Hoque K. Int J Chem Stud 2020; 8 (02) 55-61
  • 5 Sundararaman S. et al. Eng Rep 2025; 7 (04) e70099
  • 6 Rápó E, Tonk S. MDPI
  • 7 dos Santos RMM. et al. Appl Clay Sci 2017; 140: 132-139
  • 8 Ahmad A, Rafatullah M, Sulaiman O, Ibrahim MH, Hashim R. J Hazard Mater 2009; 170 (01) 357-365
  • 9 Hambisa AA, Regasa MB, Ejigu HG, Senbeto CB. Appl Water Sci 2023; 13 (01)
  • 10 de Souza PR. Environ Monit Assess 2020; 192 (03)
  • 11 Gürses A, Açıkyıldız M, Güneş K, Gürses MS. 2016: 31-45
  • 12 Bai YN. et al. J Hazard Mater 2020; 388
  • 13 Liu Y, Huang Y, Xiao A, Qiu H, Liu L. Nano 2019; 9 (01)
  • 14 Roy U. et al. 2018: 253-280
  • 15 Mahmood S, Khalid A, Arshad M, Mahmood T, Crowley DE. Crit Rev Biotechnol 2016; 36: 639
  • 16 Banerjee S, Chattopadhyaya MC. Arab J Chem 2017; 10: S1629-S1638
  • 17 Pan L, Wang Z, Yang Q, Huang R. Nanomaterials 2018; 8 (11)
  • 18 Baskar S. et al. Engineering 2025; 104161
  • 19 Sundararaman S. et al. Mater Sci Eng B 2025; 315: 118093
  • 20 Karthikeyan C. et al. J Hazard Mater 2021; 411: 124884
  • 21 Shahnawaz Khan M, Khalid M, Shahid M. What Triggers Dye Adsorption by Metal Organic Frameworks? The Current Perspectives. Royal Society of Chemistry; 2020.
  • 22 Syed A, Kumar G, Tonu NT, Chakrabarty S, Mahiuddin MD, Hoque K. Int J Chem Stud 2020; 8 (02) 55-61
  • 23 Sundararaman S. et al. Engineering 2025; 104851
  • 24 Hu H, Xu K. In: High-Risk Pollutants in Wastewater. Elsevier; 2019: 169-207
  • 25 Tan KB, Vakili M, Horri BA, Poh PE, Abdullah AZ, Salamatinia B. Sep Purif Technol 2015; 150: 229
  • 26 Nageeb M. Organic Pollutants - Monitoring, Risk and Treatment. InTech; 2013.
  • 27 Kausar A. et al. J Mol Liq 2018; 256: 395-407
  • 28 Li W, Mu B, Yang Y. Bioresour Technol 2019; 277: 157
  • 29 Afroze S, Sen TK. Water Air Soil Pollut. 2018 229. 07
  • 30 Meka U, Kumar JA, Sivamani S. S Afr J Chem Eng. 2025
  • 31 Akpomie KG, Conradie J. Sci Rep 2020; 10 (01)
  • 32 Rezaei Kalantry R, Jafari AJ, Esrafili A, Kakavandi B, Gholizadeh A, Azari A. Water Treat 2016; 57 (14) 6411-6422
  • 33 Soni R, Bhardwaj S, Shukla DP. Inorganic Pollutants in Water. Elsevier; 2020: 273-295
  • 34 Wong S, Yac NAN, Ngadi N, Hassan O, Inuwa IM. Chin J Chem Eng 2018; 26 (04) 870-878
  • 35 Saravanan P, Josephraj J, Pushpa Thillainayagam B. Environ Nanotechnol, Monit Manag 2021; 16: 100560
  • 36 Koyuncu H, Kul AR. Appl Water Sci. 2020 10. 02
  • 37 Abdel-Ghani NT, Hegazy AK, El-Chaghaby GA. Int J Environ Sci Tech 2009; 6 (02) 243-248
  • 38 Módenes AN. et al. Environ Technol 2015; 36 (22) 2892-2902
  • 39 Weber CT, Foletto EL, Meili L. Water Air Soil Pollut 2013; 224 (02)
  • 40 Fathy M, Moghny TA, Mousa MA, Abdelraheem OH, Emam AA. J Aust Ceram Soc 2020; 56 (02) 567-577
  • 41 Wang WY, Kan CW. Cellulose 2020; 27 (15) 9045-9059
  • 42 Efeovbokhan VE. et al. J Phys: Conf Ser 2019;
  • 43 Kundu R, Biswas C, Ahmed J, Naime J, Ara MH. J Chem Health Risks 2020; 10 (04) 243-252
  • 44 Cardoso NF. et al. Desalination 2011; 269 (103) 92
  • 45 Banerjee S, Chattopadhyaya MC. Arab J Chem 2017; 10: S1629-S1638
  • 46 Jawad AH, Abdulhameed AS, Mastuli MS. J Taibah Univ Sci 2020; 14 (01) 305-313
  • 47 Hegazy AK, Abdel-Ghani NT, El-Chaghaby GA. Desalin Water Treat 2011; 36 (1/3) 392-399
  • 48 Hassen JH, Ayfan AH, Joudah MT. Egypt J Chem 2022; 65 (13) 645-651
  • 49 Takam B, Acayanka E, Kamgang GY, Pedekwang MT, Laminsi S. Environ Sci Pollut Res 2017; 24 (20) 16958-16970
  • 50 Zhou Y, Lu J, Zhou Y, Liu Y. Environ Pollut
  • 51 Sharma YC. J Chem Eng Data 2010; 55 (01) 435-439
  • 52 Muthukumaran C, Sivakumar VM, Thirumarimurugan M. J Taiwan Inst Chem Eng 2016; 63: 354-362
  • 53 Arabkhani P, Asfaram A. J Hazard Mater 2020; 384
  • 54 Tanhaei B, Ayati A, Iakovleva E, Sillanpää M. Int J Biol Macromol 2020; 164: 3621-3631
  • 55 Iram M, Guo C, Guan Y, Ishfaq A, Liu H. J Hazard Mater 2010; 181 (1/3) 1039-1050
  • 56 Saxena M, Sharma N, Saxena R. Surf Interfaces 2020; 21
  • 57 Ahmad MA, Ahmad N, Bello OS. J Dispers Sci Technol 2015; 36 (05) 670-684
  • 58 Mahmoodi NM, Sadeghi U, Maleki A, Hayati B, Najafi F. J Ind Eng Chem 2014; 20 (05) 2745-2753
  • 59 Kim SH, Choi PP. Dalton Trans 2017; 46 (44) 15470-15479
  • 60 Jiang Y. et al. Carbohydr Polym 2018; 182: 106-114
  • 61 Khan EA. J Mol Liq 2018; 249: 1195-1211
  • 62 Akter N. et al. J Environ Chem Eng 2016; 4 (01) 1231-1241
  • 63 Nethaji S, Sivasamy A, Mandal AB. Int J Environ Sci Technol 2013; 10 (02) 231-242
  • 64 Gonte RR, Shelar G, Balasubramanian K. Desalin Water Treat 2014; 52 (40/42) 7797-7811
  • 65 Shikuku VO, Mishra T. Appl Water Sci 2021; 11 (06)
  • 66 Ezechi EH, Kutty SRBM, Malakahmad A, Isa MH. Process Saf Environ Prot 2015; 98: 16-32
  • 67 Laskar N, Kumar U. Int J Environ Sci Technol 2019; 16 (03) 1649-1662
  • 68 Aziam R, Chiban M, Eddaoudi H, Soudani A, Zerbet M, Sinan F. Eur Phys J: Spec Top 2017; 226 (05) 977-992
  • 69 Dhananasekaran S, Palanivel R, Pappu S. J Adv Res 2016; 7 (01) 113-124
  • 70 Dada EO, Ojo IA, Alade AO, Afolabi TJ, Amuda OS, Jameel AT. Chem Sci Int J 2020; 32-50
  • 71 El-Zahhar AA, Awwad NS, El-Katori EE. J Mol Liq 2014;
  • 72 Ghaedi M, Ghaedi AM, Negintaji E, Ansari A, Vafaei A, Rajabi M. J Ind Eng Chem 2014; 20 (04) 1793-1803
  • 73 El-Gamal SMA, Amin MS, Ahmed MA. J Environ Chem Eng 2015; 3 (03) 1702-1712
  • 74 Akpomie KG, Conradie J. Arab J Chem 2020; 13 (09) 7115-7131
  • 75 Malana MA, Ijaz S, Ashiq MN. Desalination 2010; 263 (1/3) 249-257
  • 76 Akpomie KG, Conradie J. Sci Rep 2020; 10 (01)
  • 77 Iqbal MJ, Ashiq MN. J Hazard Mater 2007; 139 (01) 57-66
  • 78 Jia Z, Li Z, Ni T, Li S. J Mol Liq 2017; 229: 285-292
  • 79 Fontana KB, Chaves ES, Sanchez JDS, Watanabe ERLR, Pietrobelli JMTA, Lenzi GG. Ecotoxicol Environ Saf 2016; 124: 329-336

Correspondence

Sagar Kumar Dutta
Chemistry, Khulna University
Khulna 9208
Bangladesh   

Publication History

Received: 22 August 2025

Accepted after revision: 29 November 2025

Article published online:
22 December 2025

© 2025. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution License, permitting unrestricted use, distribution, and reproduction so long as the original work is properly cited. (https://creativecommons.org/licenses/by/4.0/).

Georg Thieme Verlag KG
Oswald-Hesse-Straße 50, 70469 Stuttgart, Germany

Bibliographical Record
Abul Bashar, Al Shariar Hasan, Rezaul Haque, Palash Kumar Dhar, Sagar Kumar Dutta. Harnessing the Adsorption Capacity of Typha domingensis for the Sustainable Elimination of Noxious Bromophenol Blue Dyes from Aqueous Systems. Sustainability & Circularity NOW 2025; 02: a27596085.
DOI: 10.1055/a-2759-6085
  • References

  • 1 Dutta SK, Amin MK, Ahmed J, Elias M, Mahiuddin M. S Afr J Chem Eng 2022; 40: 195-208
  • 2 Sun DT. et al. ACS Cent Sci 2018; 4 (03) 349-356
  • 3 Lee A, Elam JW, Darling SB. Membrane Materials for Water Purification: Design, Development, and Application. Royal Society of Chemistry; 2016.
  • 4 Syed A, Kumar G, Tonu NT, Chakrabarty S, Mahiuddin MD, Hoque K. Int J Chem Stud 2020; 8 (02) 55-61
  • 5 Sundararaman S. et al. Eng Rep 2025; 7 (04) e70099
  • 6 Rápó E, Tonk S. MDPI
  • 7 dos Santos RMM. et al. Appl Clay Sci 2017; 140: 132-139
  • 8 Ahmad A, Rafatullah M, Sulaiman O, Ibrahim MH, Hashim R. J Hazard Mater 2009; 170 (01) 357-365
  • 9 Hambisa AA, Regasa MB, Ejigu HG, Senbeto CB. Appl Water Sci 2023; 13 (01)
  • 10 de Souza PR. Environ Monit Assess 2020; 192 (03)
  • 11 Gürses A, Açıkyıldız M, Güneş K, Gürses MS. 2016: 31-45
  • 12 Bai YN. et al. J Hazard Mater 2020; 388
  • 13 Liu Y, Huang Y, Xiao A, Qiu H, Liu L. Nano 2019; 9 (01)
  • 14 Roy U. et al. 2018: 253-280
  • 15 Mahmood S, Khalid A, Arshad M, Mahmood T, Crowley DE. Crit Rev Biotechnol 2016; 36: 639
  • 16 Banerjee S, Chattopadhyaya MC. Arab J Chem 2017; 10: S1629-S1638
  • 17 Pan L, Wang Z, Yang Q, Huang R. Nanomaterials 2018; 8 (11)
  • 18 Baskar S. et al. Engineering 2025; 104161
  • 19 Sundararaman S. et al. Mater Sci Eng B 2025; 315: 118093
  • 20 Karthikeyan C. et al. J Hazard Mater 2021; 411: 124884
  • 21 Shahnawaz Khan M, Khalid M, Shahid M. What Triggers Dye Adsorption by Metal Organic Frameworks? The Current Perspectives. Royal Society of Chemistry; 2020.
  • 22 Syed A, Kumar G, Tonu NT, Chakrabarty S, Mahiuddin MD, Hoque K. Int J Chem Stud 2020; 8 (02) 55-61
  • 23 Sundararaman S. et al. Engineering 2025; 104851
  • 24 Hu H, Xu K. In: High-Risk Pollutants in Wastewater. Elsevier; 2019: 169-207
  • 25 Tan KB, Vakili M, Horri BA, Poh PE, Abdullah AZ, Salamatinia B. Sep Purif Technol 2015; 150: 229
  • 26 Nageeb M. Organic Pollutants - Monitoring, Risk and Treatment. InTech; 2013.
  • 27 Kausar A. et al. J Mol Liq 2018; 256: 395-407
  • 28 Li W, Mu B, Yang Y. Bioresour Technol 2019; 277: 157
  • 29 Afroze S, Sen TK. Water Air Soil Pollut. 2018 229. 07
  • 30 Meka U, Kumar JA, Sivamani S. S Afr J Chem Eng. 2025
  • 31 Akpomie KG, Conradie J. Sci Rep 2020; 10 (01)
  • 32 Rezaei Kalantry R, Jafari AJ, Esrafili A, Kakavandi B, Gholizadeh A, Azari A. Water Treat 2016; 57 (14) 6411-6422
  • 33 Soni R, Bhardwaj S, Shukla DP. Inorganic Pollutants in Water. Elsevier; 2020: 273-295
  • 34 Wong S, Yac NAN, Ngadi N, Hassan O, Inuwa IM. Chin J Chem Eng 2018; 26 (04) 870-878
  • 35 Saravanan P, Josephraj J, Pushpa Thillainayagam B. Environ Nanotechnol, Monit Manag 2021; 16: 100560
  • 36 Koyuncu H, Kul AR. Appl Water Sci. 2020 10. 02
  • 37 Abdel-Ghani NT, Hegazy AK, El-Chaghaby GA. Int J Environ Sci Tech 2009; 6 (02) 243-248
  • 38 Módenes AN. et al. Environ Technol 2015; 36 (22) 2892-2902
  • 39 Weber CT, Foletto EL, Meili L. Water Air Soil Pollut 2013; 224 (02)
  • 40 Fathy M, Moghny TA, Mousa MA, Abdelraheem OH, Emam AA. J Aust Ceram Soc 2020; 56 (02) 567-577
  • 41 Wang WY, Kan CW. Cellulose 2020; 27 (15) 9045-9059
  • 42 Efeovbokhan VE. et al. J Phys: Conf Ser 2019;
  • 43 Kundu R, Biswas C, Ahmed J, Naime J, Ara MH. J Chem Health Risks 2020; 10 (04) 243-252
  • 44 Cardoso NF. et al. Desalination 2011; 269 (103) 92
  • 45 Banerjee S, Chattopadhyaya MC. Arab J Chem 2017; 10: S1629-S1638
  • 46 Jawad AH, Abdulhameed AS, Mastuli MS. J Taibah Univ Sci 2020; 14 (01) 305-313
  • 47 Hegazy AK, Abdel-Ghani NT, El-Chaghaby GA. Desalin Water Treat 2011; 36 (1/3) 392-399
  • 48 Hassen JH, Ayfan AH, Joudah MT. Egypt J Chem 2022; 65 (13) 645-651
  • 49 Takam B, Acayanka E, Kamgang GY, Pedekwang MT, Laminsi S. Environ Sci Pollut Res 2017; 24 (20) 16958-16970
  • 50 Zhou Y, Lu J, Zhou Y, Liu Y. Environ Pollut
  • 51 Sharma YC. J Chem Eng Data 2010; 55 (01) 435-439
  • 52 Muthukumaran C, Sivakumar VM, Thirumarimurugan M. J Taiwan Inst Chem Eng 2016; 63: 354-362
  • 53 Arabkhani P, Asfaram A. J Hazard Mater 2020; 384
  • 54 Tanhaei B, Ayati A, Iakovleva E, Sillanpää M. Int J Biol Macromol 2020; 164: 3621-3631
  • 55 Iram M, Guo C, Guan Y, Ishfaq A, Liu H. J Hazard Mater 2010; 181 (1/3) 1039-1050
  • 56 Saxena M, Sharma N, Saxena R. Surf Interfaces 2020; 21
  • 57 Ahmad MA, Ahmad N, Bello OS. J Dispers Sci Technol 2015; 36 (05) 670-684
  • 58 Mahmoodi NM, Sadeghi U, Maleki A, Hayati B, Najafi F. J Ind Eng Chem 2014; 20 (05) 2745-2753
  • 59 Kim SH, Choi PP. Dalton Trans 2017; 46 (44) 15470-15479
  • 60 Jiang Y. et al. Carbohydr Polym 2018; 182: 106-114
  • 61 Khan EA. J Mol Liq 2018; 249: 1195-1211
  • 62 Akter N. et al. J Environ Chem Eng 2016; 4 (01) 1231-1241
  • 63 Nethaji S, Sivasamy A, Mandal AB. Int J Environ Sci Technol 2013; 10 (02) 231-242
  • 64 Gonte RR, Shelar G, Balasubramanian K. Desalin Water Treat 2014; 52 (40/42) 7797-7811
  • 65 Shikuku VO, Mishra T. Appl Water Sci 2021; 11 (06)
  • 66 Ezechi EH, Kutty SRBM, Malakahmad A, Isa MH. Process Saf Environ Prot 2015; 98: 16-32
  • 67 Laskar N, Kumar U. Int J Environ Sci Technol 2019; 16 (03) 1649-1662
  • 68 Aziam R, Chiban M, Eddaoudi H, Soudani A, Zerbet M, Sinan F. Eur Phys J: Spec Top 2017; 226 (05) 977-992
  • 69 Dhananasekaran S, Palanivel R, Pappu S. J Adv Res 2016; 7 (01) 113-124
  • 70 Dada EO, Ojo IA, Alade AO, Afolabi TJ, Amuda OS, Jameel AT. Chem Sci Int J 2020; 32-50
  • 71 El-Zahhar AA, Awwad NS, El-Katori EE. J Mol Liq 2014;
  • 72 Ghaedi M, Ghaedi AM, Negintaji E, Ansari A, Vafaei A, Rajabi M. J Ind Eng Chem 2014; 20 (04) 1793-1803
  • 73 El-Gamal SMA, Amin MS, Ahmed MA. J Environ Chem Eng 2015; 3 (03) 1702-1712
  • 74 Akpomie KG, Conradie J. Arab J Chem 2020; 13 (09) 7115-7131
  • 75 Malana MA, Ijaz S, Ashiq MN. Desalination 2010; 263 (1/3) 249-257
  • 76 Akpomie KG, Conradie J. Sci Rep 2020; 10 (01)
  • 77 Iqbal MJ, Ashiq MN. J Hazard Mater 2007; 139 (01) 57-66
  • 78 Jia Z, Li Z, Ni T, Li S. J Mol Liq 2017; 229: 285-292
  • 79 Fontana KB, Chaves ES, Sanchez JDS, Watanabe ERLR, Pietrobelli JMTA, Lenzi GG. Ecotoxicol Environ Saf 2016; 124: 329-336

Zoom
Fig. 1 Preparation of dry TDLP.
Zoom
Fig. 2 Combination FTIR spectrum of TDLP before and after adsorption.
Zoom
Fig. 3 Determining the point of zero charge (pHzpc) of TDLP.
Zoom
Fig. 4 SEM images of (a) TDLP before adsorption and (b) TDLP after adsorption.
Zoom
Fig. 5 EDS images of (a) TDLP before adsorption and (b) TDLP after adsorption.
Zoom
Fig. 6 pH Effect of BPB dye removal.
Zoom
Fig. 7 Adsorbent dosage effect on BPB dye removal.
Zoom
Fig. 8 Contact time and initial dye concentration effect on BPB dye removal.
Zoom
Fig. 9 Pseudo-first-order kinetics at different concentrations for BPB.
Zoom
Fig. 10 Pseudo-second-order kinetics at different concentrations for BPB.
Zoom
Fig. 11 Langmuir adsorption isotherm for the TDLP.
Zoom
Fig. 12 Freundlich adsorption isotherm for the TDLP.
Zoom
Fig. 13 Plot of ln Kd vs. 1/T for evaluation of thermodynamics parameters for the adsorption of BPB onto TDLP.
Zoom
Fig. 14 Reusability Study for the TDLP.