Next Article in Journal
A Non-Stationarity Analysis of Annual Maximum Floods: A Case Study of Campaspe River Basin, Australia
Previous Article in Journal
The Impact of Aquaculture Cooperation Organization Support on Fish Farmers’ Selected Good Aquaculture Practices: Based on a Survey Data of 586 Fish Farmers in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Single-Step Modification of Brewer’s Spent Grains Using Phosphoric Acid and Application in Cheese Whey Remediation via Liquid-Phase Adsorption

by
Luiz Eduardo Nochi Castro
1,2,
Larissa Resende Matheus
1,
Rosana Rabelo Mançano
1,
William Gustavo Sganzerla
2,*,
Rafael Gabriel da Rosa
2,
Tiago Linhares Cruz Tabosa Barroso
2,
Vanessa Cosme Ferreira
2 and
Leda Maria Saragiotto Colpini
1
1
Advanced Campus of Jandaia do Sul, Federal University of Parana, 426 Dr. João Maximiano St., Jandaia do Sul 86900-000, Brazil
2
School of Food Engineering, University of Campinas, 80 Monteiro Lobato St., Campinas 13083-862, Brazil
*
Author to whom correspondence should be addressed.
Water 2023, 15(20), 3682; https://doi.org/10.3390/w15203682
Submission received: 15 September 2023 / Revised: 16 October 2023 / Accepted: 17 October 2023 / Published: 21 October 2023
(This article belongs to the Special Issue Water Use in Processing Industry)

Abstract

:
Brewer’s spent grains (BSG) are a significant by-product of beer production, and its improper disposal poses environmental challenges. This study investigated the use of BSG for activated carbon production with phosphoric acid as a chemical activator and its application in cheese whey remediation through liquid-phase adsorption. The adsorbent was thoroughly characterized through using techniques such as FTIR, SEM, N2 isotherms, and surface charge distribution. The adsorbent exhibited substantial pores, a high surface area (605.1 m2 g–1), good porosity, and positive surface charges that facilitated favorable interactions with cheese whey compounds. Equilibrium was achieved in 330 min for lactose, BOD5, and COD. The maximum adsorption capacities were 12.77 g g–1 for lactose, 3940.99 mg O2 g–1 for BOD5, and 12,857.92 mg O2 g−1 for COD at 318 K. Removing these adsorbates from cheese whey effluent reduces its organic load, enabling water reuse in the manufacturing unit, depending on its intended use. The adsorption process was spontaneous and endothermic, with ΔH° ≥ 265.72 kJ mol−1. Additionally, the activated carbon produced demonstrated impressive regeneration capability with sodium hydroxide, maintaining 75% of its adsorption capacity. These results emphasize the potential of activated carbon as an effective adsorbent for cheese whey remediation, providing a sustainable solution for waste management in the dairy industry and water reuse.

1. Introduction

Lignocellulosic resources, obtained from either natural origins or through chemical and biotechnological methods, represent sustainable source materials. There is an increasing enthusiasm for employing these resources as alternatives to fossil-based carbon in the production of various chemicals and premium biomaterials [1]. The major constituents of lignocellulosic materials include cellulose, hemicellulose, holocellulose, and lignin, and their makeup can differ depending on factors such as the source material, harvesting methods, and growth conditions [2].
Brewer’s spent grains (BSG) are a good example of lignocellulosic material. Their composition of lignin, cellulose, and hemicellulose can vary between 11 and 13%, 19 and 21%, and 34 and 48%, respectively [3,4,5]. The most significative by-product generated from brewing is BSG, estimated to be approximately 3 million tons per year [6,7]. The large amount of by-products generated can become a significant problem if not disposed of correctly. While a considerable amount is used for animal manufacturing, the quantity generated still exceeds consumption [8].
Therefore, it is necessary to find alternatives uses for BSG, such as in a biorefinery context in the production of sugars and amino acids [9], bioenergy [10], and ethanol [11]. Additionally, research indicates that BSG is a very promising material for producing activated carbon for use in adsorption processes [12,13,14,15]. BSG possesses great characteristics for use as an adsorbent due to its high concentration of carbon, silicon, and aluminum, as well as its high volatile content, which facilitates pore formation [16].
Several methods can be used to activate carbonaceous materials, with the majority of used methods being chemical methods involving phosphoric acid (H3PO4) and potassium hydroxide (KOH) [3]. As previously mentioned, various studies have employed phosphoric acid (H3PO4) as an activating agent. For instance, from banana peel, Romero-Ayana et al. [17] obtained activated carbon with a surface area over 2000 m2 g−1 and a yield higher than 35%. Highlighting the significance of method selection, it is crucial to consider factors such as the time required for synthesis, the nature and potential harm of the reagents, the properties of the resulting material, and, most importantly, the level of complexity and the equipment needed [18,19,20].
Cheese whey (CW) is the primary wastewater from cheese manufacturing, constituting a residue containing fats, proteins, and carbohydrates, resulting in a high organic load [21,22]. This effluent exhibits high resistance in terms of COD (700–78,000 mg L−1) and BOD (500–17,000 mg L−1) [22,23,24]. Numerous methods and strategies have been explored for managing the wastewater produced by the cheese manufacturing industry. These encompass a range of approaches, such as physical–chemical and biological methods, electrochemical solutions, constructed wetlands, advanced oxidation processes, and hybrid systems, among others [25,26,27].
Nguyen et al. [28] demonstrated that activated carbon possesses the capability to eliminate stubborn organic compounds. The utilization of waste in the form of activated carbon as part of the adsorption process has emerged as an alternative to conventional waste treatment procedures. This approach has demonstrated effectiveness, is cost-effective, eliminates the need for chemical processes, produces no sludge, and proves highly efficient in both discoloration and degradation processes [29]. Adsorption using activated carbon has become an emerging technology that the industry can explore for the treatment of liquid waste, and several studies have shown that the use of adsorption with activated carbon in industrial effluents yields better results when compared to conventional methods [30,31]. Wastewater treatment contributes to water savings by purifying polluted water for reuse in various industrial and commercial applications. This not only reduces the need for fresh water but also helps companies comply with stringent regulations, save on operational costs, preserve natural water resources, prevent environmental pollution, and promote sustainable water management [32,33].
The objective of this work was to evaluate the use of BSG to produce activated carbon using H3PO4 as a chemical activator and apply it in the remediation of cheese whey via liquid-phase adsorption.

2. Materials and Methods

2.1. Materials

The BSG samples were provided by craft beer producers in the northern region of Paraná State, Brazil. Before use, the BSG were quartered, and opposite quarters were used for analysis, while the others were separated. Subsequently, the samples were washed to remove any physical contaminants, such as leaves and sticks. Additionally, some of the BSG were dried in a forced circulation oven (Lucadema, model 82/480) at 60 °C for 24 h to remove all moisture from the samples. The CW was provided by a cheese factory in the northern region of Paraná State, Brazil, and the CW was obtained from the production of “Minas Frescal” cheese on site. The CW was extracted during the curd cutting stage of producing coagulated milk. After the collection, the CW was filtered with a cheesecloth to remove any particulate matter from the production process.

2.2. Preparation of Adsorbent

This study employed a one-step acid activation procedure using H3PO4 [18]. In this process, 20 g BSG were combined with 20 g H3PO4 solution (85% w/v). The resulting mixture was stirred for 5 min and allowed to sit undisturbed for 8 h. Afterward, the material was rinsed with distilled water to eliminate any excess acid and subsequently dried at 105 °C overnight. Following the drying step, the sample underwent carbonization at 400 °C for 4 h, with a temperature increase rate of 10 °C min−1 within a muffle furnace (Solidsteel, model SSFM 16L). The sample was then subjected to washing with a NaHCO3 solution (2% w/v) until it reached a pH of 7. Finally, the sample was dried for approximately 4 h at 120 °C and coded as ACPO4. In order to gauge the effectiveness of the synthesized adsorbent, a commercial activated carbon (Øparticle = 5 mm) was employed as a reference sample in all tests.

2.3. Adsorbent Characterization

The samples were characterized using scanning electron microscopy (SEM) with energy dispersive spectroscopy (VEGA3—TESCAN) to obtain micrographs and determine their elemental composition. N2 adsorption/desorption isotherm measurements were carried out using a sorption analyzer (NOVA 2000e—Quantachrome Instruments). Fourier-transform infrared spectroscopy (FTIR, Spectrum Two—Perkin Elmer) was also used. The point of zero charges (PZC) or surface charge distribution was determined by mixing 50 mg adsorbent with 50 mL aqueous solution at different initial pH conditions (2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12), adjusting the solutions with HCl or NaOH 0.1 mol L−1, and measuring the pH after 24 h of equilibrium [30].

2.4. Cheese Whey Adsorption Assays

To evaluate the capacity of the adsorbent in real effluent decontamination processes, adsorption tests with cheese whey were carried out. The parameters evaluated during the adsorption tests included the reduction in lactose content, reduction in biochemical oxygen demand (BOD5), and reduction in chemical oxygen demand (COD). These parameters were evaluated to determine the level of decontamination of the actual effluent. Lactose content was determined via titration using the adapted Lane–Eynon method [34], while the BOD5 and COD contents were determined using the American Public Health Association method [35].
The acid modification of the BSG was evaluated through a test comparing the biomass and the synthesized material to commercial activated carbon to estimate the impact of the activation method on the adsorbent’s adsorptive capacity. The cheese whey volume was 50 mL at pH = 6.5, the adsorbent concentration was 10 g L−1 (m = 0.05 g), the temperature was 25 °C, and the reaction time was 24 h. The initial experiments were conducted by altering the activated carbon concentration to identify the ideal quantity of adsorbent. This allowed for subsequent tests to be performed with the minimal amount of material required, thus optimizing its utilization. The concentration of the ACPO4 activated carbon was varied from 0.5 g L−1 to 10 g L−1 (m = 0.025–0.5 g), and we used 50 mL of cheese whey at pH = 6.5, 25 °C, under agitation in a shaker (TE-424 Tecnal) for 12 h.
The equations for calculating the percentage removal of lactose, BOD5, and COD (R, %) (Equation (1)), adsorption capacity at any time for the lactose (qL, g g−1) (Equation (2)), and adsorption capacity for the BOD5 and COD (qB, qC, mg g−1) (Equation (3)) are presented below:
R = C 0   C C 0 100
q L = ( x 0 x t )   ρ   m   V
q B =   q C = y 0 y t m   V
where R (%) is the percentage of reduction of the analyzed parameter, C0 (g of lactose 100 g−1 of cheese whey or mg of O2 L−1 of cheese whey) is the initial concentration of the parameters lactose, BOD5, and COD, respectively, C (g of lactose 100 g−1 of cheese whey or mg of O2 L−1 of cheese whey) is the final concentration of the parameters lactose, BOD5, and COD, qL (g of lactose g−1 of adsorbent) is the quantity of lactose adsorbed at time t, xo (g of lactose g−1 of cheese whey) is the initial concentration of lactose, xt (g of lactose g−1 of cheese whey) is the time t concentration of lactose, ρ (1023.2 g of cheese whey L−1 of cheese whey) is the specific mass of cheese whey, m (g) is the mass of the adsorbent, V (L) is the volume of the solution of cheese whey, qB and qC (mg of O2 g−1 of adsorbent) represent the quantity of biological/chemical oxygen dissolved in the CW at time t, yo (mg of O2 g−1 of cheese whey) is the initial quantity of biological/chemical oxygen dissolved in the CW, and yt (mg of O2 g−1 of cheese whey) is the time t quantity of biological/chemical oxygen dissolved in the CW.

2.4.1. Kinetic Modeling

To determine the equilibrium time (te) for the adsorption assays with the CW, a kinetic study was performed using the same parameters as the preliminary tests. Aliquots were collected every 15 min for the first 2 h and then every hour until equilibrium.
The theoretical amount of lactose, BOD5, and COD adsorbed at equilibrium (qe) was determined using four kinetic models: pseudo-first-order (PFO), pseudo-second-order (PSO), and Avrami fractional-order (AFO) [18,36,37,38]. These values were then compared with the experimental values obtained for the parameters adsorbed at equilibrium (qeExp). The pseudo-first-order (Equation (4)), pseudo-second-order (Equation (5)), and Avrami fractional-order (Equation (6)) models are represented below:
q t = q e ( 1 e k 1 t )
q t = k 2 q e 2   t   1 + k 2 q e t
q t = q e ( 1 e ( k A t ) n A )
where qe (mg g−1) is the quantity of the parameters adsorbed at equilibrium, k1 (min−1) represents the pseudo-first-order kinetic constant, t (min) represents the reaction time, k2 (g mg−1 min−1) represents the pseudo-second-order kinetic constant, kA (min−1) represents the Avrami kinetic constant, and nA (dimensionless) is the exponent Avrami of time.

2.4.2. Equilibrium Modeling

The present study involved delving into the adsorption equilibrium process and specifically examining the relationship between the concentration of parameters (Ce) in the liquid phase (solution) and the concentration of parameters (qe) on the surface of the solid phase (adsorbent). This investigation involved analyzing adsorption isotherms and assessing how temperature impacts the maximum adsorbent capacity of the materials. Experimental isotherm data were collected at temperatures of 288, 298, 308, and 318 K and then adjusted using four models: Langmuir (Equation (7)), Freundlich (Equation (8)), Dubinin–Radushkevich (Equation (9, 10 and 11)), and Hill (Equation (12)) [39,40,41].
q   e = q L K L C e 1 + K L C e
q e = K F C e 1 n
q e =   q mDR exp β ε 2
ε = RTln C S C e
E s = 1 2 β
q   e = n H q mH 1 + C 1 / 2 C e n H
where qL (mg g−1) is the maximum biosorption capacity of the Langmuir model, KL (L mg−1) is the Langmuir constant, KF ((mg g−1) (mg L−1)−1/n) is the Freundlich constant, 1/n (dimensionless) is the heterogeneity factor, β (mol2 kJ−2) is the Dubinin–Radushkevich constant, ε is the Polanyi potential, R (8314 J mol−1 K−1) is the universal gas constant, CS (mg L−1) is the solubility, C1/2 (mg L−1) is the concentration at half saturation, nH (dimensionless) is the number of molecules per site, and qmH (mg g−1) is the density of receptor sites.

2.4.3. Thermodynamic Modeling

A thermodynamic study was carried out to improve our understanding of the energetic changes involved during the adsorption process at different temperatures. Enthalpy (ΔH°), entropy (ΔS°), the free energy of adsorption (ΔG°), and activation energy (Ea) were calculated for a better understanding of the effects of the parameters under the interaction mechanism between the adsorbent and adsorbate involved in the adsorption process. The equations of Van ’t Hoff, Gibbs–Helmholtz, and Arrhenius [42,43] were used to determine the parameters mentioned above. The standard Gibbs free energy change (ΔG°, kJ mol−1), enthalpy change (ΔH°, kJ mol−1), entropy change (ΔS°, kJ mol−1 K−1), and activation energy (Ea, kJ mol−1) were calculated using the following Equations:
K e = K M W γ C W Γ
G ° = RTln K e
G ° = H °   T S °
ln K e = S ° R H ° RT
k = A   e   E a RT
ln   ( k ) = ln   A   E a RT
where Ke (dimensionless) is the equilibrium constant, K (L mg−1) is the constant parameter from the most suitable isotherm fit, MW (26,600 g mol−1) is the molecular weight of the solid constituents in cheese whey, γ (dimensionless, assuming γCW = 1) is the activity coefficient of cheese whey, Γ (1 mol L−1) is the unitary activity coefficient of cheese whey, T (K) is the temperature, R (8.314 × 10−3 kJ mol−1 K−1) is the universal gas constant, ∆S° (kJ mol−1 K−1) represents adsorption entropy, ∆H° (kJ mol−1) represents adsorption enthalpy, ∆G° (kJ mol−1) represents the free energy of adsorption, k’(min−1) is the constant parameter from the most suitable kinetic fit, A (dimensionless) is a pre-exponential factor, and Ea (kJ mol−1) represents the activation energy.

2.5. Desorption and Regeneration Experiments

After the adsorption of CW was carried out under the optimal conditions (see conditions in Section 3.4), desorption experiments of CW from ACPO4 were carried out using HCl (0.1 mol L−1) and NaOH (0.1 mol L−1) solutions as eluents [44,45,46]. A total of 500 mg of ACPO4 loaded with CW were in contact with 50 mL of the eluent (HCl or NaOH) and agitated for 2 h at 200 rpm for desorption. The concentration of lactose, BOD5, and COD in the liquid phase were determined as in Section 2.4. The regeneration assays were carried out 5 times.

2.6. Statistical Analysis

Nonlinear techniques were applied to assess the compatibility of the kinetic and equilibrium data through the utilization of the Simplex method and the Levenberg–Marquardt algorithm. These methods were employed within the fitting capabilities of the Microcal Origin 2021 software. To gauge the appropriateness of the kinetic and equilibrium models, various metrics were employed, including the residual sum of squares (RSS), determination coefficient (R2), adjusted determination coefficient (R2adj), standard deviation of residues (SD), and the Bayesian information criterion (BIC). Mathematical representations for these metrics are presented in Equations (19)–(23).
RSS = i n q i ,   exp q i ,   model 2
R 2 = i n q i , exp q ¯ exp 2 i n q i ,   exp q i ,   model 2 i n q i , exp q ¯ exp 2
R adj 2 = 1 1 R 2 . n 1 n p 1
SD = 1 n p . i n q i ,   exp q i ,   model 2
BIC = nLn RSS n + pLn ( n )
In the equations provided above, each qi, model represents the predicted theoretical q value for a specific individual as per the model’s prediction. Correspondingly, qi, exp stands for the individual experimental q value obtained through actual experimentation. The symbol q ¯ exp denotes the mean of all the measured experimental q values. The variable n signifies the total count of experiments conducted, while p represents the count of parameters in the fitting model.
Our analysis included the presentation of R2adj, SD, and BIC values to facilitate a comparison between the various kinetics and equilibrium models outlined in this study. The ideal model exhibited R2adj values nearing 1.000, lower SD values, and minimized BIC values. However, selecting the optimal kinetic and equilibrium model involves a more nuanced assessment beyond relying solely on R2 values, especially when these models encompass differing parameter quantities. Hence, it becomes imperative to ascertain whether enhancements in R2 values stem from an increase in parameters or if, in a physical sense, the model featuring a greater number of parameters more effectively elucidates the underlying process.
Nevertheless, the disparity in BIC values among models could be decisive when the discrepancy in BIC values is ≤2.0, indicating the absence of a significant distinction between the two models. Within the range of BIC value differences spanning 2 to 6, a favorable trend arises toward the model possessing the lower BIC value, signifying its enhanced suitability. In instances where the range of BIC value differences stretches from 6 to 10, a robust likelihood emerges that the model with the lower BIC value constitutes the most fitting choice. However, if the contrast in BIC values attains ≥ 10.0, a confident prediction can be made that the model endowed with the lower BIC value is unequivocally the superior fit.

3. Results and Discussion

3.1. Adsorbents Characterization

Figure 1 shows the SEM micrographs obtained for the materials used in this study. All the materials presented a rugged surface with laminar and granular particles, which are characteristic of biomass and activated carbons obtained from biomass [47,48,49]. The ACPO4 (Figure 1e,f) displayed a great distribution of pores on the surface, characteristic of materials with excellent adsorptive capacity and high surface area, as shown in previous publications [50,51]. These favorable characteristics can allow for the favorable penetration of cheese whey molecules into the pores of the particles [52].
The elemental composition of the materials was determined via EDS analysis, as shown in Table 1. Overall, the primary element in all materials was carbon, indicating that the materials were carbonized throughout the synthesis process, as was the case for ACPO4 and the commercial activated carbon. This high carbon content might also be a sign of the adsorptive capacity of the adsorbents, as shown in the literature [18,53]. Because the materials were carbonized in the presence of air, oxygen was also expected. Silicon, magnesium, potassium, and calcium emanated from the inherent composition of the barley that generated the BSG during beer production [16,54]. The presence of phosphorus in the ACPO4 comes from the reagent used during synthesis (H3PO4).
Figure 2 shows the N2 adsorption–desorption isotherms for the adsorbents obtained.
In Figure 2a,b, the isotherms are like Type I isotherms, typical of microporous materials, while the isotherms in Figure 2c resemble Type II isotherms, characteristic of microporous and mesoporous materials [55,56]. This shift in the isotherm format occurred for the chemically activated material, where the volume of N2 adsorbed decreased, probably due to the interaction between the base material and chemical/thermal reactions that occurred during the activation process, modifying the surface structure of the materials, especially the shape and depth of the pores. Other publications in the literature have reported similar behavior regarding the shift in isotherm format and adsorptive volume [57,58]. The distribution of the pores in the materials can be seen in Figure 2d. The range of the pore size distribution was concentrated around 1.8 nm, which is regarded as a micropore range and is frequently observed in works that use activated carbon as an adsorbent [59,60,61].
Table 2 shows the BET’s surface area (So), mean pore diameter (dp), and pore volumes (Vp) of the different materials analyzed in this study. It was possible to observe that after the BSG was activated with H3PO4, the surface area value increased almost six-fold. This is possible because chemical activation with this acid is dehydrating. In this mechanism, the acid attacks the porous structure of the material and, due to its high affinity with water, it removes the hydrogen and oxygen atoms from the base material, dehydrating it, causing the porous structure to improve and, consequently, its surface area and pore volume to increase. This increase in surface area after activation shows that this method managed to improve the surface characteristics of the base material. It is also worth noting that the ACPO4 had a higher surface area than the commercial activated carbon that was used as our control, demonstrating that this synthesis is efficient and can produce high-quality adsorbent materials. By observing the pore diameter values, it was possible to see that the activated carbons had pore diameters characteristic of micropore materials (dp < 2 nm), while the BSG had a slightly higher value, falling in the mesoporous material range (2 ≤ dp ≤ 50 nm). These findings align with the outcomes observed in the isotherm graphs depicted in Figure 2 and are consistent with data from previous research studies, as shown in Table 3.
The acquired FTIR spectra for the adsorbents are shown in Figure 3. Two major regions were detected in the BSG spectrum (Figure 3a), with bands between 7500 and 7000 cm−1 in the first region that are characteristic vibration signals of the CH3, CH2, and CH bonds in the second overtone region, which may indicate the possible presence of lignocellulosic compounds such as cellulose, hemicellulose, and lignin [64,65]. The bands between 5400 and 4500 cm−1 in the second region made it possible to identify the vibration characteristics between the C-C and H2O bonds in the combination region [8,66]. For the commercial activated carbon sample (Figure 3b), the same band regions of the BSG were found, indicating the presence of carbonaceous structures in the material.
In Figure 3c, the spectrum is divided into three regions. The first and second regions are like those in Figure 3a,b; however, the third region with bands between 4710 and 4210 cm−1 with a maximum at approximately 4333 cm−1 is characteristic of the combination between the C-H + C-H groups (methyl, methylene, or methine), which also indicates that the material contains a carbon structure [67].
Figure 4 displays the outcomes of the PZC analysis performed on the materials. The PZC assesses the surface characteristics of the adsorbent, providing insights into whether the material tends to be more acidic or basic in nature. When the pHPZC exceeds the pH of the solution, it suggests a negative charge on the material’s surface, whereas when the pHPZC is lower than the pH of the solution, it indicates a positive surface charge [18]. The pH of the cheese whey used during the adsorption tests was pH = 6.5.
Our analysis revealed a variety of pHPZC values for the adsorbents. Specifically, for both the commercial activated carbon and BSG, their pHPZC values exceeded the pH of the cheese whey, indicating a negatively charged surface on these materials. In contrast, ACPO4 exhibited a pHPZC lower than the pH of the cheese whey, signifying a positively charged surface [68,69]. It is known that the electrical charges of cheese whey proteins have a negative charge, so the physisorption process of whey molecules will be favored onto materials with a positive surface charge, such as ACPO4 [70,71,72].

3.2. Cheese Whey Characterization

Before the adsorption assays, the complete characterization of the cheese whey was carried out to determine the initial parameters of the effluent, and the results for lactose, BOD5, and COD were 4.79 g of lactose 100 g−1 of cheese whey, 4329.60 mg of O2 L−1 of cheese whey, and 72,578.30 mg of O2 L−1 of cheese whey, respectively. It was possible to observe that the COD/BOD5 ratio was approximately 17, which indicates that this effluent (COD/BOD5 > 4) has a very low level of biodegradability, meaning that the use of physicochemical methods would be required for the degradation of this pollutant, such as adsorption [73,74]. Furthermore, it is known that due to its natural composition, cheese whey is a highly recalcitrant pollutant, especially due to its solid content (mostly lactose) and its high turbidity, which prevents the passage of light and hinders the oxygenation of water bodies.
Thus, it is necessary to monitor these parameters as indicators of the degradation of this pollutant [75,76].

3.3. Adsorption Experiment of Cheese Whey onto Activated Carbons

3.3.1. Acid Modification Effect

Figure 5 shows the results of the acid modification test on the BSG. It was evident that acid modification had a positive impact on the removal of lactose, increasing from 17.9% when using 40.6% of BSG when using the commercial adsorbent. Similarly, BOD5 removal improved from 12.5% with BSG to 29.7% with the commercial adsorbent, and COD removal increased from 21.5% with BSG to 59.4% with the commercial adsorbent. This outcome was anticipated because biomass (BSG) naturally possesses a low adsorption capacity due to its lack of surface porosity and limited interaction with the adsorbate [77]. This differs from other materials (e.g., commercial 5 mm activated carbon and ACPO4) that have undergone activation through physical or chemical means [78]. Such modifications aim to enhance their structural, morphological, and surface characteristics, allowing the adsorbents to acquire adsorption capacity either through increased surface area, increased porosity, or more favorable surface interactions [79].
ACPO4 exhibited the highest removal percentages for lactose (67.6%), BOD5 (47.3%), and COD (88.5%) when compared to biomass (BSG) and the commercial adsorbent (Commercial 5 mm). This superior performance can be attributed to the intrinsic characteristics of the material (outlined in Section 3.1), such as its higher specific surface area (So) (605.1 m2 g−1) and positive surface charge distribution (pHPCZ). These characteristics enhanced the adsorption of molecules found in the CW, particularly in conjunction with the extensive pore distribution observed on the surface of the adsorbent, as revealed via SEM analysis. As a result, ACPO4 was chosen as the preferred adsorbent for the subsequent adsorption experiments involving CW in this study.

3.3.2. ACPO4 Dosage Effect

The impact of ACPO4 dosage on CW adsorption is depicted in Figure 6. The removal of lactose (R, %) increased as the ACPO4 dosage changed from 0.5 to 10 g L−1 (Figure 6a), rising from 25.2% to 63.3%. Similarly, for BOD5 (Figure 6b), it increased from 20.7% to 45.9%, and for COD (Figure 6c), it increased from 72.2% to 91.1%. This phenomenon can be explained by the higher amount of adsorbent added to the reaction mixture. When there is a greater mass of adsorbent present on the material’s surface, it leads to an increased number of active sites becoming accessible, consequently causing a natural increase in the removal of adsorbates from the CW [80,81]. Table 4 presents a comparison between the results obtained by the adsorption technique (in this study) in the removal of lactose, BOD5, and COD from cheese whey with other studies in the literature using various techniques for wastewater treatment. It becomes clear that the adsorption technique is as effective for removing organic content from cheese whey as the different techniques used in other studies in the literature. Thus, the use of liquid-phase adsorption is a viable alternative for treating effluents from dairy facilities, reducing their organic load and allowing this waste to be used as residual water for cleaning the external areas of dairy facilities, as feedwater in boiler systems for steam and electricity generation, or for irrigating various crops.
In addition, it is essential to note that the adsorption capacity is inversely proportional to the mass of the adsorbent present in the medium. Consequently, the adsorption capacity, in the case of lactose uptake, decreased from 41.34 g lactose g−1 adsorbent with a dosage of 0.5 g L−1 to 5.20 g lactose g−1 adsorbent with 10 g L−1 of ACPO4. Similarly, for BOD5 uptake, it decreased from 400.93 mg O2 g−1 adsorbent using 0.5 g L−1 of ACPO4 to 35.96 mg O2 g−1 using 10 g L−1, and for COD uptake, it decreased from 5252.04 mg O2 g−1 adsorbent using 0.5 g L−1 of ACPO4 to 400.41 mg O2 g−1 using 10 g L−1. Consequently, the subsequent experimental procedures for obtaining kinetic profiles and the equilibrium studies were conducted using a dosage of 2 g L−1. This dosage was chosen as it represents the intersection of the curves obtained in the dosage experiments [86,87].

3.3.3. ACPO4 Adsorption Kinetics

The kinetics of lactose, BOD5, and COD uptake onto ACPO4 were investigated using three different models: PFO, PSO, and AFO models (Table 5).
The statistical evaluation of the kinetic models considered R2adj, SD, and BIC values (Section 2.6). The model with the highest R2adj (close to 1), lowest SD, and lowest BIC values for all the studied adsorbates was the AFO model, indicating that this model was the most suitable for fitting the dispersion of the kinetic data. The Bayesian information criterion becomes more significant when ΔBIC ≥ 2 [88,89]. When ΔBIC ≥ 10, the model with the lowest BIC score is undoubtedly the best statistically fitted model [90]. For lactose, the ΔBIC value between PFO and AFO was 11.84, and between PSO and AFO, it was 14.41. For BOD5, the ΔBIC value between PFO and AFO was 11.41, and between PSO and AFO, it was 12.24. Finally, for COD, the ΔBIC value between PFO and AFO was 29.09, and between PSO and AFO, it was 48.93. Hence, based on the BIC values presented in Table 4, it is evident that AFO is the most suitable kinetic model for characterizing the adsorption kinetics of lactose, BOD5, and COD onto ACPO4.
Because different kinetic adsorption models utilize different units for the kinetic rate constant (k), making direct comparisons regarding the time needed to reach equilibrium becomes challenging. Therefore, alternative measures like t1/2 (the time to achieve 50% saturation of the adsorbent) and t0.95 (the time to achieve 95% saturation of the adsorbent) were employed. These values were determined by interpolation from the adjusted curve (as shown in Table 5). For the AFO kinetic model, the t1/2 values ranged from 24.86 to 73.88 min, while the t0.95 values ranged from 200.34 to 315.99 min. Since the AFO model was deemed the most suitable for describing the kinetics, these values offer a more accurate representation of the time-related parameter. To ensure that the system reached equilibrium, it was necessary for the contact time to exceed t0.95. Consequently, the decision was made to use contact times of 330 min to ensure equilibrium attainment.
Figure 7 illustrates the AFO model for all the adsorbates studied in the uptake of CW onto ACPO4.

3.3.4. ACPO4 Adsorption Equilibrium

The adsorption isotherms for lactose, BOD5, and COD onto ACPO4 were determined under optimized conditions using Langmuir, Freundlich, Dubinin–Radushkevich, and Hill models at temperatures of 288 K, 298 K, 308 K, and 318 K. The isotherm parameters are shown in Table 6.
The equilibrium models were assessed based on the R2adj, SD, and BIC values. As mentioned earlier, ΔBIC values between Langmuir, Freundlich, Dubinin–Radushkevich, and Hill models were all ≤ 10 at temperatures ranging from 288 K to 318 K (Table 5). This suggests a strong likelihood that the model with the lowest BIC value is the most appropriate choice. In this case, the Hill model proved to be the most suitable for describing the adsorption of lactose, BOD5, and COD onto ACPO4 [91,92].
It is also worth noting that the Hill isotherms obtained through data fitting are favorable (Figure 8), indicating a high adsorption capacity while utilizing a minimal quantity of ACPO4 material [85,86]. With an increase in Ce, the adsorption capacity also increased significantly, reaching values of approximately 5.7 g g−1 for lactose, 1150 mg O2 g−1 for BOD5, and 11,000 mg O2 g−1 for COD (experimental values). It is evident that the adsorption capacity benefitted from higher temperatures, with the highest values observed at 318 K. This phenomenon can be attributed to the thermal agitation effect, where an increase in agitation rate reduces the film resistance to mass transfer around the adsorbent particles, enhancing the adsorption of molecules present in the cheese whey onto the active sites of ACPO4 [93,94]. In summary, these results confirm that ACPO4 can effectively be employed for the treatment of cheese whey through the adsorption process. This reduces its organic load and facilitates the remediation of this pollutant, allowing for potential reuse as treated water or its safe return to the aquatic environment.

3.3.5. ACPO4 Thermodynamic Parameters

The thermodynamic equilibrium parameters were evaluated using the Van ’t Hoff equation, where the Langmuir parameter (KL, expressed in L mg−1) was utilized to estimate the thermodynamic equilibrium constant (Ke) [43,95,96]. The thermodynamic data are presented in Table 7.
At first, the decline in ΔG° with an increase in temperature implies that the adsorption of lactose, BOD5, and COD onto ACPO4 became increasingly favorable at elevated temperatures. Additionally, the consistently negative ΔG° values at all temperature levels indicate that the adsorption process is spontaneous and favorable. ΔG° values between −20 and 0 kJ mol−1 indicate a predominance of the physisorption phenomenon, as observed at the lower temperatures in this study (288–298 K), ranging from −1.51 to −6.26 kJ mol−1. However, at higher temperatures (308–318 K), the ΔG° values are ≤−20 kJ mol−1, ranging from −20.44 to −26.99 kJ mol−1, indicating a shift in the predominant adsorption equilibrium mechanism [97,98]. Additionally, positive ΔS° values for all adsorbates imply some structural adjustments at the interface during the adsorption process.
Moreover, the positive ΔH° values provide confirmation that the adsorption of these adsorbates is an endothermic process. This observation aligns with the fact that as temperature rises, the adsorption of lactose, BOD5, and COD onto ACPO4 increases in line with Le Chatelier’s principle, which pertains to perturbing chemical equilibrium [99]. Moreover, the ΔH° values exceeding 40 kJ mol−1 for all adsorbates suggest that the process is not solely driven by physisorption. In the case of lactose, BOD5, and COD adsorption, it may involve a combination of physisorption and ion exchange or chemisorption onto the surface of the ACPO4 adsorbent [42,100].
Finally, it is worth noting that the activation energy (Ea) values associated with the adsorption of lactose, BOD5, and COD onto ACPO4 exhibited remarkable similarity, differing by only about 6 kJ mol−1. This slight variation in activation energy could indicate a subtle difference in the affinity between the adsorbate and the adsorbent’s surface. However, the closely matched activation energy values suggest that the adherence process for lactose, BOD5, and COD is quite similar. The magnitude of the activation energy provides valuable insights into the fundamental mechanism of the adsorption process. Usually, physisorption, which is based on physical attraction, is associated with activation energies below 40 kJ mol−1. On the other hand, chemisorption, a chemical process, requires higher energy levels, typically those exceeding 40 kJ mol−1. The activation energy (Ea) values we determined for the adsorption of lactose, BOD5, and COD indicate that physisorption is the prevailing mechanism governing the adsorption of these substances on the ACPO4 surface [101,102,103]. This fact is also supported by the results listed in Table 6. When observing the value of Es obtained through the Dubinin–Radushkevich model, one can infer about the adsorption mechanism. For Es < 8 (kJ mol−1), the process is mainly based on physical adsorption, which can be observed for all the values obtained for the adsorption of lactose, BOD5, and COD. This reaffirms that indeed a physisorption process was the governing mechanism in this study [104].

3.4. Proposal of the Adsorption Mechanism

The proposed adsorption mechanism was based on findings from the characterization results of ACPO4, including FTIR and pHPZC, as well as insights from thermodynamic modeling. The adsorbent was anticipated to be a lignocellulosic substance primarily composed of cellulose and lignin, featuring functional groups like CO, CC, and OH- [105,106]. The pH level of the CW’s operation was 6.5, resulting in a positive charge on the material’s surface at this pH. Regarding the molecules found in the CW, it was established that proteins presented in the CW carry a negative charge, which promotes the bonding between the adsorbate and adsorbent. Finally, thermodynamic information revealed that the primary adsorption mechanism for CW on ACPO4 is physisorption, driven by physical interactions such as hydrogen bonding, π–π interactions, anion–π interactions, and electrostatic interactions, as depicted in Figure 9 [89,100,107].

3.5. Regeneration and Reuse of the ACPO4 Adsorbent

Firstly, adsorption experiments were conducted under specific conditions, including a 2 g L−1 dosage of the adsorbent, an initial pH of 6.5, a temperature of 318 K, and a contact time of 330 min. Consequently, the adsorption capacity for lactose reached approximately 3.50 g lactose g−1 adsorbent; for BOD5, it was around 500.00 mg O2 g−1 adsorbent, and for COD, it was approximately 5300.00 mg O2 g−1 adsorbent. Following this, HCl and NaOH were evaluated as eluents, with 0.1 mol L−1 NaOH proving to be the most effective eluent, releasing all adsorbates into the liquid phase within 75 min. These adsorption–desorption cycles were repeated five times, and the outcomes are illustrated in Figure 10.

4. Conclusions

The remediation of cheese whey using activated carbon synthesized through single-step acid modification with H3PO4 (ACPO4) via liquid-phase adsorption has proven to be an efficient method for treating and recovering this residue from the dairy industry. ACPO4 exhibits physical and chemical characteristics favorable for the adsorption process, including a high surface area of 605.1 m2 g−1, well-developed pores with high volume and diameter, and positive surface charges that facilitate favorable electrostatic interactions with the molecules present in cheese whey. The kinetic data confirm that equilibrium was reached after 330 min for the parameters evaluated in the adsorption process (lactose, BOD5, and COD). The isothermal curves follow the Hill isotherm model, with a maximum adsorption capacity of 12.77 g g−1 for lactose, 3940.99 mg O2 g−1 for BOD5, and 12,857.92 mg O2 g−1 for COD at the highest temperature (318 K). The thermodynamic results indicate that adsorption (ΔH° ≥ 265.72 kJ mol−1) for all the adsorbates was spontaneous, favorable, and endothermic. ACPO4 can be regenerated with NaOH and reused up to four times, with only a 25% reduction in its adsorption capacity. In conclusion, based on these findings, it can be stated that using ACPO4 as an adsorbent for cheese whey remediation is promising and offers an alternative for managing dairy industry residues and reusing water resources. By allowing the water recovered after adsorptive processes to be used within the industrial environment for various purposes, significant water savings are generated.

Author Contributions

Conceptualization, L.E.N.C.; methodology, L.E.N.C., L.R.M. and R.R.M.; investigation, L.E.N.C., W.G.S., R.G.d.R., T.L.C.T.B. and V.C.F.; writing—original draft preparation, L.E.N.C., L.R.M., R.R.M., R.G.d.R., T.L.C.T.B. and V.C.F.; writing—review and editing, W.G.S. and L.M.S.C.; supervision, L.M.S.C.; project administration, L.M.S.C.; funding acquisition, L.M.S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was not supported by any dedicated grants from public, commercial, or nonprofit organizations.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to express their gratitude to CAPES for providing financial support for this research article.

Conflicts of Interest

The authors affirm that they have no identifiable competing financial interests or personal affiliations that might have influenced the research presented in this paper.

References

  1. Tišma, M.; Bucic-Kojic, A.; Planinic, M. Bio-Based Products from Lignocellulosic Waste Biomass: A State of the Art. Chem. Biochem. Eng. Q. 2021, 35, 139–156. [Google Scholar] [CrossRef]
  2. Zeko-Pivač, A.; Tišma, M.; Žnidaršič-Plazl, P.; Kulisic, B.; Sakellaris, G.; Hao, J.; Planinić, M. The Potential of Brewer’s Spent Grain in the Circular Bioeconomy: State of the Art and Future Perspectives. Front. Bioeng. Biotechnol. 2022, 10, 870744. [Google Scholar] [CrossRef]
  3. Osman, A.I.; O’Connor, E.; McSpadden, G.; Abu-Dahrieh, J.K.; Farrell, C.; Al-Muhtaseb, A.H.; Harrison, J.; Rooney, D.W. Upcycling Brewer’s Spent Grain Waste into Activated Carbon and Carbon Nanotubes for Energy and Other Applications via Two-Stage Activation. J. Chem. Technol. Biotechnol. 2020, 95, 183–195. [Google Scholar] [CrossRef]
  4. Mahmood, A.S.N.; Brammer, J.G.; Hornung, A.; Steele, A.; Poulston, S. The Intermediate Pyrolysis and Catalytic Steam Reforming of Brewers Spent Grain. J. Anal. Appl. Pyrolysis 2013, 103, 328–342. [Google Scholar] [CrossRef]
  5. Vanreppelen, K.; Vanderheyden, S.; Kuppens, T.; Schreurs, S.; Yperman, J.; Carleer, R. Activated Carbon from Pyrolysis of Brewer’s Spent Grain: Production and Adsorption Properties. Waste Manag. Res. 2014, 32, 634–645. [Google Scholar] [CrossRef]
  6. Sganzerla, W.G.; Ampese, L.C.; Mussatto, S.I.; Forster-Carneiro, T. A Bibliometric Analysis on Potential Uses of Brewer’s Spent Grains in a Biorefinery for the Circular Economy Transition of the Beer Industry. Biofuels Bioprod. Biorefining 2021, 15, 1965–1988. [Google Scholar] [CrossRef]
  7. Sganzerla, W.G.; Buller, L.S.; Mussatto, S.I.; Forster-Carneiro, T. Techno-Economic Assessment of Bioenergy and Fertilizer Production by Anaerobic Digestion of Brewer’s Spent Grains in a Biorefinery Concept. J. Clean. Prod. 2021, 297, 126600. [Google Scholar] [CrossRef]
  8. Castro, L.E.N.; Mançano, R.R.; Battocchio, D.A.J.; Colpini, L.M.S. Adsorção de Alimento Corante Usando Carvão Ativado de Cervejarias’ Grãos Gastos. Acta Scientiarum. Technol. 2022, 45, e60443. [Google Scholar] [CrossRef]
  9. Sganzerla, W.G.; Viganó, J.; Castro, L.E.N.; Maciel-Silva, F.W.; Rostagno, M.A.; Mussatto, S.I.; Forster-Carneiro, T. Recovery of Sugars and Amino Acids from Brewers’ Spent Grains Using Subcritical Water Hydrolysis in a Single and Two Sequential Semi-Continuous Flow-through Reactors. Food Res. Int. 2022, 157, 111470. [Google Scholar] [CrossRef]
  10. Dudek, M.; Świechowski, K.; Manczarski, P.; Koziel, J.A.; Białowiec, A. The Effect of Biochar Addition on the Biogas Production Kinetics from the Anaerobic Digestion of Brewers’ Spent Grain. Energies 2019, 12, 1518. [Google Scholar] [CrossRef]
  11. Rojas-Chamorro, J.A.; Cara, C.; Romero, I.; Ruiz, E.; Romero-García, J.M.; Mussatto, S.I.; Castro, E. Ethanol Production from Brewers’ Spent Grain Pretreated by Dilute Phosphoric Acid. Energy Fuels 2018, 32, 5226–5233. [Google Scholar] [CrossRef]
  12. Wierzba, S.; Rajfur, M.; Nabrdalik, M.; Kłos, A. Assessment of the Influence of Counter Ions on Biosorption of Copper Cations in Brewer’s Spent Grain—Waste Product Generated during Beer Brewing Process. Microchem. J. 2019, 145, 196–203. [Google Scholar] [CrossRef]
  13. Jackowski, M.; Niedźwiecki, Ł.; Jagiełło, K.; Uchańska, O.; Trusek, A. Brewer’s Spent Grains—Valuable Beer Industry By-Product. Biomolecules 2020, 10, 1669. [Google Scholar] [CrossRef]
  14. Wierzba, S.; Kłos, A. Heavy Metal Sorption in Biosorbents—Using Spent Grain from the Brewing Industry. J. Clean. Prod. 2019, 225, 112–120. [Google Scholar] [CrossRef]
  15. de Araújo, T.P.; Tavares, F.d.O.; Vareschini, D.T.; Barros, M.A.S.D. Biosorption Mechanisms of Cationic and Anionic Dyes in a Low-Cost Residue from Brewer’s Spent Grain. Environ. Technol. 2020, 42, 2925–2940. [Google Scholar] [CrossRef]
  16. Castro, L.E.N.; Colpini, L.M.S. All-around Characterization of Brewers’ Spent Grain. Eur. Food Res. Technol. 2021, 247, 3013–3021. [Google Scholar] [CrossRef]
  17. Romero-Anaya, A.J.; Molina, A.; Garcia, P.; Ruiz-Colorado, A.A.; Linares-Solano, A.; Salinas-Martínez de Lecea, C. Phosphoric Acid Activation of Recalcitrant Biomass Originated in Ethanol Production from Banana Plants. Biomass Bioenergy 2011, 35, 1196–1204. [Google Scholar] [CrossRef]
  18. de Castro, L.E.N.; Battocchio, D.A.J.; Ribeiro, L.F.; Colpini, L.M.S. Development of Adsorbent Materials Using Residue from Coffee Industry and Application in Food Dye Adsorption Processes. Braz. Arch. Biol. Technol. 2022, 66, e23210125. [Google Scholar] [CrossRef]
  19. Li, M.; Xiao, R. Preparation of a Dual Pore Structure Activated Carbon from Rice Husk Char as an Adsorbent for CO2 Capture. Fuel Process. Technol. 2019, 186, 35–39. [Google Scholar] [CrossRef]
  20. Ahmed, M.B.; Hasan Johir, M.A.; Zhou, J.L.; Ngo, H.H.; Nghiem, L.D.; Richardson, C.; Moni, M.A.; Bryant, M.R. Activated Carbon Preparation from Biomass Feedstock: Clean Production and Carbon Dioxide Adsorption. J. Clean. Prod. 2019, 225, 405–413. [Google Scholar] [CrossRef]
  21. Venetsaneas, N.; Antonopoulou, G.; Stamatelatou, K.; Kornaros, M.; Lyberatos, G. Using Cheese Whey for Hydrogen and Methane Generation in a Two-Stage Continuous Process with Alternative PH Controlling Approaches. Bioresour. Technol. 2009, 100, 3713–3717. [Google Scholar] [CrossRef] [PubMed]
  22. Aghili, F.; Ghoreyshi, A.A.; Rahimpour, A.; Rahimnejad, M. Dynamic Behavior of the Adsorption, Activated Sludge and Combined Activated Sludge-Adsorption Process for Treatment of Cheese Whey Wastewater. Desalination Water Treat. 2015, 57, 16404–16414. [Google Scholar] [CrossRef]
  23. Martins, R.C.; Quinta-Ferreira, R.M. Final Remediation of Post-Biological Treated Milk Whey Wastewater by Ozone. Int. J. Chem. React. Eng. 2010, 8, 142. [Google Scholar] [CrossRef]
  24. Carvalho, F.; Prazeres, A.R.; Rivas, J. Cheese Whey Wastewater: Characterization and Treatment. Sci. Total Environ. 2013, 445–446, 385–396. [Google Scholar] [CrossRef] [PubMed]
  25. Tatoulis, T.I.; Tekerlekopoulou, A.G.; Akratos, C.S.; Pavlou, S.; Vayenas, D.V. Aerobic Biological Treatment of Second Cheese Whey in Suspended and Attached Growth Reactors. J. Chem. Technol. Biotechnol. 2015, 90, 2040–2049. [Google Scholar] [CrossRef]
  26. Chatzipaschali, A.A.; Stamatis, A.G. Biotechnological Utilization with a Focus on Anaerobic Treatment of Cheese Whey: Current Status and Prospects. Energies 2012, 5, 3492–3525. [Google Scholar] [CrossRef]
  27. Kotoulas, A.; Agathou, D.; Triantaphyllidou, I.E.; Tatoulis, T.I.; Akratos, C.S.; Tekerlekopoulou, A.G.; Vayenas, D.V. Zeolite as a Potential Medium for Ammonium Recovery and Second Cheese Whey Treatment. Water 2019, 11, 136. [Google Scholar] [CrossRef]
  28. Nguyen, L.N.; Hai, F.I.; Kang, J.; Price, W.E.; Nghiem, L.D. Removal of Trace Organic Contaminants by a Membrane Bioreactor–Granular Activated Carbon (MBR–GAC) System. Bioresour. Technol. 2012, 113, 169–173. [Google Scholar] [CrossRef]
  29. Barakan, S.; Aghazadeh, V. The Advantages of Clay Mineral Modification Methods for Enhancing Adsorption Efficiency in Wastewater Treatment: A Review. Environ. Sci. Pollut. Res. 2020, 28, 2572–2599. [Google Scholar] [CrossRef]
  30. Castro, L.E.N.; Santos, J.V.F.; Fagnani, K.C.; Alves, H.J.; Colpini, L.M.S. Evaluation of the Effect of Different Treatment Methods on Sugarcane Vinasse Remediation. J. Environ. Sci. Health 2019, 54, 791–800. [Google Scholar] [CrossRef]
  31. Azmi, N.S.; Yunos, K.F.M. Wastewater Treatment of Palm Oil Mill Effluent (POME) by Ultrafiltration Membrane Separation Technique Coupled with Adsorption Treatment as Pre-Treatment. Agric. Agric. Sci. Procedia 2014, 2, 257–264. [Google Scholar] [CrossRef]
  32. Rashid, R.; Shafiq, I.; Akhter, P.; Iqbal, M.J.; Hussain, M. A State-of-the-Art Review on Wastewater Treatment Techniques: The Effectiveness of Adsorption Method. Environ. Sci. Pollut. Res. 2021, 28, 9050–9066. [Google Scholar] [CrossRef] [PubMed]
  33. Al-Hazmi, H.E.; Mohammadi, A.; Hejna, A.; Majtacz, J.; Esmaeili, A.; Habibzadeh, S.; Saeb, M.R.; Badawi, M.; Lima, E.C.; Mąkinia, J. Wastewater Reuse in Agriculture: Prospects and Challenges. Environ. Res. 2023, 236, 116711. [Google Scholar] [CrossRef]
  34. da Silva, L.C.A.; Lafetá Junior, J.A.Q.; Leite, M.O.; Fontes, E.A.F.; Coimbra, J.S.R. Comparative Appraisal of HPLC, Chloramine-T and Lane–Eynon Methods for Quantification of Carbohydrates in Concentrated Dairy Products. Int. J. Dairy. Technol. 2020, 73, 795–800. [Google Scholar] [CrossRef]
  35. Baird, R.; Rice, E.; Eaton, A. Standard Methods for the Examination of Water and Wastewater, 23rd ed.; Association, W.A.P.H., Ed.; American Water Works Association: Denver, CO, USA; Water Environment Federation: Alexandria, VA, USA, 2017. [Google Scholar]
  36. Lagergreen, S. Zur Theorie Der Sogenannten Adsorption Gelöster Stoffe. Z. Chem. Ind. Kolloide 1907, 2, 15. [Google Scholar] [CrossRef]
  37. Ho, Y.S.; McKay, G. Sorption of Dye from Aqueous Solution by Peat. Chem. Eng. J. 1998, 70, 115–124. [Google Scholar] [CrossRef]
  38. Cestari, A.R.; Vieira, E.F.S.; Lopes, E.C.N.; Da Silva, R.G. Kinetics and Equilibrium Parameters of Hg(II) Adsorption on Silica–Dithizone. J. Colloid. Interface Sci. 2004, 272, 271–276. [Google Scholar] [CrossRef]
  39. Langmuir, I. The Constitution and Fundamental Properties of Solids and Liquids. Part I. Solids. J. Am. Chem. Soc. 1916, 38, 2221–2295. [Google Scholar] [CrossRef]
  40. Nguyen, C.; Do, D.D. The Dubinin–Radushkevich Equation and the Underlying Microscopic Adsorption Description. Carbon 2001, 39, 1327–1336. [Google Scholar] [CrossRef]
  41. Wakkel, M.; Khiari, B.; Zagrouba, F. Textile Wastewater Treatment by Agro-Industrial Waste: Equilibrium Modelling, Thermodynamics and Mass Transfer Mechanisms of Cationic Dyes Adsorption onto Low-Cost Lignocellulosic Adsorbent. J. Taiwan. Inst. Chem. Eng. 2019, 96, 439–452. [Google Scholar] [CrossRef]
  42. Hameed, B.H.; Ahmad, A.A.; Aziz, N. Isotherms, Kinetics and Thermodynamics of Acid Dye Adsorption on Activated Palm Ash. Chem. Eng. J. 2007, 133, 195–203. [Google Scholar] [CrossRef]
  43. Lima, E.C.; Hosseini-Bandegharaei, A.; Moreno-Piraján, J.C.; Anastopoulos, I. A Critical Review of the Estimation of the Thermodynamic Parameters on Adsorption Equilibria. Wrong Use of Equilibrium Constant in the Van’t Hoof Equation for Calculation of Thermodynamic Parameters of Adsorption. J. Mol. Liq. 2019, 273, 425–434. [Google Scholar] [CrossRef]
  44. Li, C.; Song, X.; Hein, S.; Wang, K. The Separation of GMP from Milk Whey Using the Modified Chitosan Beads. Adsorption 2010, 16, 85–91. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Campbell, R.; Drake, M.A.; Zhong, Q. Decolorization of Cheddar Cheese Whey by Activated Carbon. J. Dairy Sci. 2015, 98, 2982–2991. [Google Scholar] [CrossRef] [PubMed]
  46. Guliyev, R.; Akgün, M.; Sayın Börekçi, B.; Sadak, O.; Esen, Y. Modelling and Process Optimization of Cheese Whey Wastewater Treatment Using Magnetic Nanoparticles. Biomass Convers. Biorefin 2022, 1, 1–12. [Google Scholar] [CrossRef]
  47. Alrowais, R.; Said, N.; Bashir, M.T.; Ghazy, A.; Alwushayh, B.; Daiem, M.M.A. Adsorption of Diphenolic Acid from Contaminated Water onto Commercial and Prepared Activated Carbons from Wheat Straw. Water 2023, 15, 555. [Google Scholar] [CrossRef]
  48. Peixoto, B.S.; Mota, L.S.D.O.; Oliveira, P.C.O.D.; Veloso, M.C.D.C.; Romeiro, G.A.; Moraes, M.C.D. Highly Functionalized Microporous Activated Biochar from Syagrus Coronata Waste: Production, Characterization, and Application in Adsorption Studies. Water 2022, 14, 3525. [Google Scholar] [CrossRef]
  49. Ameen, M.M.; Moustafa, A.A.; Mofeed, J.; Hasnaoui, M.; Olanrewaju, O.S.; Lazzaro, U.; Guerriero, G. Factors Affecting Efficiency of Biosorption of Fe (III) and Zn (II) by Ulva Lactuca and Corallina Officinalis and Their Activated Carbons. Water 2021, 13, 3421. [Google Scholar] [CrossRef]
  50. Kifetew, M.; Alemayehu, E.; Fito, J.; Worku, Z.; Prabhu, S.V.; Lennartz, B. Adsorptive Removal of Reactive Yellow 145 Dye from Textile Industry Effluent Using Teff Straw Activated Carbon: Optimization Using Central Composite Design. Water 2023, 15, 1281. [Google Scholar] [CrossRef]
  51. Li, Y.; Zhang, J.; Liu, H. Removal of Chloramphenicol from Aqueous Solution Using Low-Cost Activated Carbon Prepared from Typha Orientalis. Water 2018, 10, 351. [Google Scholar] [CrossRef]
  52. Bevilacqua, R.C.; Preigschadt, I.A.; Netto, M.S.; Georgin, J.; Franco, D.S.P.; Mallmann, E.S.; Silva, L.F.O.; Pinto, D.; Foletto, E.L.; Dotto, G.L. One Step Acid Modification of the Residual Bark from Campomanesia Guazumifolia Using H2SO4 and Application in the Removal of 2,4-Dichlorophenoxyacetic from Aqueous Solution. J. Environ. Sci. Health Part B 2021, 56, 995–1006. [Google Scholar] [CrossRef]
  53. Quan, C.; Zhou, Y.; Wu, C.; Xu, G.; Feng, D.; Zhang, Y.; Gao, N. Valorization of Solid Digestate into Activated Carbon and Its Potential for CO2 Capture. J. Anal. Appl. Pyrolysis 2023, 169, 105874. [Google Scholar] [CrossRef]
  54. Gil, A.; Pallarés, J.; Arauzo, I.; Cortés, C. Pyrolysis and CO2 Gasification of Barley Straw: Effect of Particle Size Distribution and Chemical Composition. Powder Technol. 2023, 424, 118539. [Google Scholar] [CrossRef]
  55. Lawtae, P.; Phothong, K.; Tangsathitkulchai, C.; Wongkoblap, A. Analysis of Gas Adsorption and Pore Development of Microporous-Mesoporous Activated Carbon Based on GCMC Simulation and a Surface Defect Model. J. Porous Mater. 2023, 1, 1–17. [Google Scholar] [CrossRef]
  56. Zheng, Y.; Li, Q.; Yu, G.; Jiang, B.; Ren, B.; Wang, S. Characteristics of in Situ Synthesized Activated Carbon and Metal-Organic Framework Composites for CH4/N2 Gas Mixture Separation. Greenh. Gases Sci. Technol. 2023, 13, 67–80. [Google Scholar] [CrossRef]
  57. Alothman, Z.A. A Review: Fundamental Aspects of Silicate Mesoporous Materials. Materials 2012, 5, 2874–2902. [Google Scholar] [CrossRef]
  58. Benedetti, V.; Patuzzi, F.; Baratieri, M. Characterization of Char from Biomass Gasification and Its Similarities with Activated Carbon in Adsorption Applications. Appl. Energy 2018, 227, 92–99. [Google Scholar] [CrossRef]
  59. Serafin, J.; Dziejarski, B.; Cruz Junior, O.F.; Sreńscek-Nazzal, J. Design of Highly Microporous Activated Carbons Based on Walnut Shell Biomass for H2 and CO2 Storage. Carbon 2023, 201, 633–647. [Google Scholar] [CrossRef]
  60. Cruz, O.F.; Campello-Gómez, I.; Casco, M.E.; Serafin, J.; Silvestre-Albero, J.; Martínez-Escandell, M.; Hotza, D.; Rambo, C.R. Enhanced CO2 Capture by Cupuassu Shell-Derived Activated Carbon with High Microporous Volume. Carbon. Lett. 2023, 33, 727–735. [Google Scholar] [CrossRef]
  61. Cruz, O.F.; Gómez, I.C.; Rodríguez-Reinoso, F.; Silvestre-Albero, J.; Rambo, C.R.; Martínez-Escandell, M. Activated Carbons with High Micropore Volume Obtained from Polyurethane Foams for Enhanced CO2 Adsorption. Chem. Eng. Sci. 2023, 273, 118671. [Google Scholar] [CrossRef]
  62. Zięzio, M.; Charmas, B.; Jedynak, K.; Hawryluk, M.; Kucio, K. Preparation and Characterization of Activated Carbons Obtained from the Waste Materials Impregnated with Phosphoric Acid(V). Appl. Nanosci. 2020, 10, 4703–4716. [Google Scholar] [CrossRef]
  63. Wu, H.; Chen, R.; Du, H.; Zhang, J.; Shi, L.; Qin, Y.; Yue, L.; Wang, J. Synthesis of Activated Carbon from Peanut Shell as Dye Adsorbents for Wastewater Treatment. Adsorpt. Sci. Technol. 2019, 37, 34–48. [Google Scholar] [CrossRef]
  64. McLellan, T.M.; Aber, J.D.; Martin, M.E.; Melillo, J.M.; Nadelhoffer, K.J. Determination of Nitrogen, Lignin, and Cellulose Content of Decomposing Leaf Material by near Infrared Reflectance Spectroscopy. Can. J. For. Res. 2011, 21, 1684–1688. [Google Scholar] [CrossRef]
  65. Li, X.; Sun, C.; Zhou, B.; He, Y. Determination of Hemicellulose, Cellulose and Lignin in Moso Bamboo by Near Infrared Spectroscopy. Sci. Rep. 2015, 5, 17210. [Google Scholar] [CrossRef] [PubMed]
  66. Xiaobo, Z.; Jiewen, Z.; Povey, M.J.W.; Holmes, M.; Hanpin, M. Variables Selection Methods in Near-Infrared Spectroscopy. Anal. Chim. Acta 2010, 667, 14–32. [Google Scholar] [CrossRef]
  67. Lee, S.Y.; Shim, H.E.; Yang, J.E.; Choi, Y.J.; Jeon, J. Continuous Flow Removal of Anionic Dyes in Water by Chitosan-Functionalized Iron Oxide Nanoparticles Incorporated in a Dextran Gel Column. Nanomaterials 2019, 9, 1164. [Google Scholar] [CrossRef]
  68. Castro, L.E.N.; Matheus, L.R.; Albuquerque, L.J.C.; Gasparini, L.J.; Fagnani, K.C.; Alves, H.J.; Colpini, L.M.S. Production of Nanostructured Crystalline Composite Using Residual Ashes from Flocculated Sludge Burning Process in a Poultry Slaughterhouse Wastewater Treatment System. Cerâmica 2022, 68, 427–440. [Google Scholar] [CrossRef]
  69. de Castro, L.E.N.; Meurer, E.C.; Alves, H.J.; dos Santos, M.A.R.; Vasques, E.d.C.; Colpini, L.M.S. Photocatalytic Degradation of Textile Dye Orange-122 Via Electrospray Mass Spectrometry. Braz. Arch. Biol. Technol. 2020, 63, e20180573. [Google Scholar] [CrossRef]
  70. Chen, G.Q.; Qu, Y.; Gras, S.L.; Kentish, S.E. Separation Technologies for Whey Protein Fractionation. Food Eng. Rev. 2023, 15, 438–465. [Google Scholar] [CrossRef]
  71. Yogarathinam, L.T.; Gangasalam, A.; Ismail, A.F.; Arumugam, S.; Narayanan, A. Concentration of Whey Protein from Cheese Whey Effluent Using Ultrafiltration by Combination of Hydrophilic Metal Oxides and Hydrophobic Polymer. J. Chem. Technol. Biotechnol. 2018, 93, 2576–2591. [Google Scholar] [CrossRef]
  72. Arunkumar, A.; Etzel, M.R. Milk Protein Concentration Using Negatively Charged Ultrafiltration Membranes. Foods 2018, 7, 134. [Google Scholar] [CrossRef]
  73. Von Sperling, M. Introdução à Qualidade Das Águas e Ao Tratamento de Esgotos, 2nd ed.; Universidade Federal de Minas Gerais: Belo Horizonte, Brazil, 1996; ISBN 9788570411143. [Google Scholar]
  74. Hu, R.; Liu, Y.; Zhu, G.; Chen, C.; Hantoko, D.; Yan, M. COD Removal of Wastewater from Hydrothermal Carbonization of Food Waste: Using Coagulation Combined Activated Carbon Adsorption. J. Water Process Eng. 2022, 45, 102462. [Google Scholar] [CrossRef]
  75. Gonzalez-Piedra, S.; Hernández-García, H.; Perez-Morales, J.M.; Acosta-Domínguez, L.; Bastidas-Oyanedel, J.R.; Hernandez-Martinez, E. A Study on the Feasibility of Anaerobic Co-Digestion of Raw Cheese Whey with Coffee Pulp Residues. Energies 2021, 14, 3611. [Google Scholar] [CrossRef]
  76. Bella, K.; Venkateswara Rao, P. Anaerobic Co-Digestion of Cheese Whey and Septage: Effect of Substrate and Inoculum on Biogas Production. J. Environ. Manag. 2022, 308, 114581. [Google Scholar] [CrossRef]
  77. Wen, J.; Liu, Z.; Xi, H.; Huang, B. Synthesis of Hierarchical Porous Carbon with High Surface Area by Chemical Activation of (NH4)2C2O4 Modified Hydrochar for Chlorobenzene Adsorption. J. Environ. Sci. 2023, 126, 123–137. [Google Scholar] [CrossRef]
  78. Tang, Z.; Gao, J.; Zhang, Y.; Du, Q.; Feng, D.; Dong, H.; Peng, Y.; Zhang, T.; Xie, M. Ultra-Microporous Biochar-Based Carbon Adsorbents by a Facile Chemical Activation Strategy for High-Performance CO2 Adsorption. Fuel Process. Technol. 2023, 241, 107613. [Google Scholar] [CrossRef]
  79. Singh, G.; Maria Ruban, A.; Geng, X.; Vinu, A. Recognizing the Potential of K-Salts, Apart from KOH, for Generating Porous Carbons Using Chemical Activation. Chem. Eng. J. 2023, 451, 139045. [Google Scholar] [CrossRef]
  80. Zhu, F.; Wang, Z.; Huang, J.; Hu, W.; Xie, D.; Qiao, Y. Efficient Adsorption of Ammonia on Activated Carbon from Hydrochar of Pomelo Peel at Room Temperature: Role of Chemical Components in Feedstock. J. Clean. Prod. 2023, 406, 137076. [Google Scholar] [CrossRef]
  81. Zhang, J.W.; Mariska, S.; Pap, S.; Tran, H.N.; Chao, H.P. Enhanced Separation Capacity of Carbonaceous Materials (Hydrochar, Biochar, and Activated Carbon) toward Potential Toxic Metals through Grafting Copolymerization. Sep. Purif. Technol. 2023, 320, 124229. [Google Scholar] [CrossRef]
  82. Gutiérrez, J.L.R.; Encina, P.A.G.; Fdz-Polanco, F. Anaerobic Treatment of Cheese-Production Wastewater Using a UASB Reactor. Bioresour. Technol. 1991, 37, 271–276. [Google Scholar] [CrossRef]
  83. Rodgers, M.; Zhan, X.-M.; Dolan, B. Mixing Characteristics and Whey Wastewater Treatment of a Novel Moving Anaerobic Biofilm Reactor. J. Environ. Sci. Health Part A 2004, 39, 2183–2193. [Google Scholar] [CrossRef] [PubMed]
  84. Rivas, J.; Prazeres, A.R.; Carvalho, F.; Beltrán, F. Treatment of Cheese Whey Wastewater: Combined Coagulation−Flocculation and Aerobic Biodegradation. J. Agric. Food Chem. 2010, 58, 7871–7877. [Google Scholar] [CrossRef]
  85. Rivas, J.; Prazeres, A.R.; Carvalho, F. Aerobic Biodegradation of Precoagulated Cheese Whey Wastewater. J. Agric. Food Chem. 2011, 59, 2511–2517. [Google Scholar] [CrossRef] [PubMed]
  86. Georgin, J.; Franco, D.S.P.; Schadeck Netto, M.; Allasia, D.; Foletto, E.L.; Oliveira, L.F.S.; Dotto, G.L. Transforming Shrub Waste into a High-Efficiency Adsorbent: Application of Physalis Peruvian Chalice Treated with Strong Acid to Remove the 2,4-Dichlorophenoxyacetic Acid Herbicide. J. Environ. Chem. Eng. 2021, 9, 104574. [Google Scholar] [CrossRef]
  87. Caponi, N.; Schnorr, C.; Franco, D.S.P.; Netto, M.S.; Vedovatto, F.; Tres, M.V.; Zabot, G.L.; Abaide, E.R.; Silva, L.F.O.; Dotto, G.L. Potential of Subcritical Water Hydrolyzed Soybean Husk as an Alternative Biosorbent to Uptake Basic Red 9 Dye from Aqueous Solutions. J. Environ. Chem. Eng. 2022, 10, 108603. [Google Scholar] [CrossRef]
  88. de Azevedo, C.F.; Machado, F.M.; de Souza, N.F.; Silveira, L.L.; Lima, E.C.; Andreazza, R.; Bergamnn, C.P. Comprehensive Adsorption and Spectroscopic Studies on the Interaction of Carbon Nanotubes with Diclofenac Anti-Inflammatory. Chem. Eng. J. 2023, 454, 140102. [Google Scholar] [CrossRef]
  89. Tran, H.N. Applying Linear Forms of Pseudo-Second-Order Kinetic Model for Feasibly Identifying Errors in the Initial Periods of Time-Dependent Adsorption Datasets. Water 2023, 15, 1231. [Google Scholar] [CrossRef]
  90. Guo, Y.; Su, C.; Chen, H.; Wang, J.; Liu, B.; Zeng, Z.; Li, L. Hierarchical Porous Carbon with Tunable Apertures and Nitrogen/Oxygen Heteroatoms for Efficient Adsorption and Separation of VOCs. Chem. Eng. J. 2023, 471, 144558. [Google Scholar] [CrossRef]
  91. de Azevedo, C.F.; Rodrigues, D.L.C.; Silveira, L.L.; Lima, E.C.; Osorio, A.G.; Andreazza, R.; de Pereira, C.M.P.; Poletti, T.; Machado Machado, F. Comprehensive Adsorption and Spectroscopic Studies on the Interaction of Magnetic Biochar from Black Wattle Sawdust with Beta-Blocker Metoprolol. Bioresour. Technol. 2023, 388, 129708. [Google Scholar] [CrossRef] [PubMed]
  92. Teixeira, R.A.; Thue, P.S.; Lima, É.C.; Grimm, A.; Naushad, M.; Dotto, G.L.; dos Reis, G.S. Adsorption of Omeprazole on Biobased Adsorbents Doped with Si/Mg: Kinetic, Equilibrium, and Thermodynamic Studies. Molecules 2023, 28, 4591. [Google Scholar] [CrossRef]
  93. Yanan, C.; Srour, Z.; Ali, J.; Guo, S.; Taamalli, S.; Fèvre-Nollet, V.; da Boit Martinello, K.; Georgin, J.; Franco, D.S.P.; Silva, L.F.O.; et al. Adsorption of Paracetamol and Ketoprofenon Activated Charcoal Prepared from the Residue of the Fruit of Butiacapitate: Experiments and Theoretical Interpretations. Chem. Eng. J. 2023, 454, 139943. [Google Scholar] [CrossRef]
  94. Gabriela Elvir-Padilla, L.; Ileana Mendoza-Castillo, D.; Villanueva-Mejía, F.; Bonilla-Petriciolet, A. Molecular Aggregation Effect on the Antagonistic Adsorption of Pharmaceuticals from Aqueous Solution Using Bone Char: DFT Calculations and Multicomponent Experimental Studies. J. Mol. Liq. 2023, 369, 120957. [Google Scholar] [CrossRef]
  95. Lima, E.C.; Sher, F.; Guleria, A.; Saeb, M.R.; Anastopoulos, I.; Tran, H.N.; Hosseini-Bandegharaei, A. Is One Performing the Treatment Data of Adsorption Kinetics Correctly? J. Environ. Chem. Eng. 2021, 9, 104813. [Google Scholar] [CrossRef]
  96. Lima, É.C.; Dehghani, M.H.; Guleria, A.; Sher, F.; Karri, R.R.; Dotto, G.L.; Tran, H.N. Adsorption: Fundamental Aspects and Applications of Adsorption for Effluent Treatment. Green. Technol. Defluoridation Water 2021, chapter 3, 41–88. [Google Scholar] [CrossRef]
  97. Cantu, Y.; Remes, A.; Reyna, A.; Martinez, D.; Villarreal, J.; Ramos, H.; Trevino, S.; Tamez, C.; Martinez, A.; Eubanks, T.; et al. Thermodynamics, Kinetics, and Activation Energy Studies of the Sorption of Chromium(III) and Chromium(VI) to a Mn3O4 Nanomaterial. Chem. Eng. J. 2014, 254, 374–383. [Google Scholar] [CrossRef]
  98. Dotto, G.L.; Vieillard, J.; Pinto, D.; Lütke, S.F.; Silva, L.F.O.; dos Reis, G.S.; Lima, É.C.; Franco, D.S.P. Selective Adsorption of Gadolinium from Real Leachate Using a Natural Bentonite Clay. J. Environ. Chem. Eng. 2023, 11, 109748. [Google Scholar] [CrossRef]
  99. Khadhri, N.; El Khames Saad, M.; Ben Mosbah, M.; Moussaoui, Y. Batch and Continuous Column Adsorption of Indigo Carmine onto Activated Carbon Derived from Date Palm Petiole. J. Environ. Chem. Eng. 2019, 7, 102775. [Google Scholar] [CrossRef]
  100. Rodrigues, L.A.; Maschio, L.J.; da Silva, R.E.; da Silva, M.L.C.P. Adsorption of Cr(VI) from Aqueous Solution by Hydrous Zirconium Oxide. J. Hazard. Mater. 2010, 173, 630–636. [Google Scholar] [CrossRef]
  101. Boparai, H.K.; Joseph, M.; O’Carroll, D.M. Kinetics and Thermodynamics of Cadmium Ion Removal by Adsorption onto Nano Zerovalent Iron Particles. J. Hazard. Mater. 2011, 186, 458–465. [Google Scholar] [CrossRef]
  102. Wang, H.; Wang, S.; Wang, S.; Fu, L.; Zhang, L. Efficient Metal-Organic Framework Adsorbents for Removal of Harmful Heavy Metal Pb(II) from Solution: Activation Energy and Interaction Mechanism. J. Environ. Chem. Eng. 2023, 11, 109335. [Google Scholar] [CrossRef]
  103. Song, Y.; Zhao, Z.; Li, J.; You, Y.; Ma, X.; Li, J.; Cheng, X. Preparation of Silicon-Doped Ferrihydrite for Adsorption of Lead and Cadmium: Property and Mechanism. Chin. Chem. Lett. 2021, 32, 3169–3174. [Google Scholar] [CrossRef]
  104. Cojocaru, C.; Samoila, P.; Pascariu, P. Chitosan-Based Magnetic Adsorbent for Removal of Water-Soluble Anionic Dye: Artificial Neural Network Modeling and Molecular Docking Insights. Int. J. Biol. Macromol. 2019, 123, 587–599. [Google Scholar] [CrossRef] [PubMed]
  105. Prado, J.M.; Lachos-Perez, D.; Forster-Carneiro, T.; Rostagno, M.A. Sub- and Supercritical Water Hydrolysis of Agricultural and Food Industry Residues for the Production of Fermentable Sugars: A Review. Food Bioprod. Process. 2016, 98, 95–123. [Google Scholar] [CrossRef]
  106. Abaide, E.R.; Tres, M.V.; Zabot, G.L.; Mazutti, M.A. Reasons for Processing of Rice Coproducts: Reality and Expectations. Biomass Bioenergy 2019, 120, 240–256. [Google Scholar] [CrossRef]
  107. Li, S.; Qi, B.; Luo, J.; Zhuang, Y.; Wan, Y. Ultrafast Selective Adsorption of Pretreatment Inhibitors from Lignocellulosic Hydrolysate with Metal-Organic Frameworks: Performance and Adsorption Mechanisms. Sep. Purif. Technol. 2021, 275, 119183. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs: BSG (a) 500× magnification and (b) 5000× magnification; commercial activated carbon (5 mm) (c) 500× magnification and (d) 5000× magnification; ACPO4 (e) 500× magnification and (f) 5000× magnification.
Figure 1. SEM micrographs: BSG (a) 500× magnification and (b) 5000× magnification; commercial activated carbon (5 mm) (c) 500× magnification and (d) 5000× magnification; ACPO4 (e) 500× magnification and (f) 5000× magnification.
Water 15 03682 g001
Figure 2. N2 measurements: (a) adsorption and desorption isotherm for BSG; (b) adsorption and desorption isotherm for the commercial activated carbon (5 mm); (c) adsorption and desorption isotherm for ACPO4; (d) pore size distribution.
Figure 2. N2 measurements: (a) adsorption and desorption isotherm for BSG; (b) adsorption and desorption isotherm for the commercial activated carbon (5 mm); (c) adsorption and desorption isotherm for ACPO4; (d) pore size distribution.
Water 15 03682 g002
Figure 3. FTIR spectra for (a) the BSG, (b) the commercial activated carbon (5 mm), and (c) ACPO4.
Figure 3. FTIR spectra for (a) the BSG, (b) the commercial activated carbon (5 mm), and (c) ACPO4.
Water 15 03682 g003
Figure 4. Point of zero charge for the adsorbents.
Figure 4. Point of zero charge for the adsorbents.
Water 15 03682 g004
Figure 5. Effect of acid modification in cheese whey adsorption.
Figure 5. Effect of acid modification in cheese whey adsorption.
Water 15 03682 g005
Figure 6. Effect of ACPO4 dosage in cheese whey adsorption: (a) lactose; (b) BOD5; (c) COD.
Figure 6. Effect of ACPO4 dosage in cheese whey adsorption: (a) lactose; (b) BOD5; (c) COD.
Water 15 03682 g006
Figure 7. AFO kinetic curves of cheese whey uptake. Conditions: adsorbent dosage of 2 g L−1, temperature of 25 °C, initial pH of 6.5.
Figure 7. AFO kinetic curves of cheese whey uptake. Conditions: adsorbent dosage of 2 g L−1, temperature of 25 °C, initial pH of 6.5.
Water 15 03682 g007
Figure 8. Adsorption isotherms of cheese whey on ACPO4 at different temperatures: (a) lactose; (b) BOD5; (c) COD. Conditions: adsorbent dosage of 2 g L−1, initial pH of 6.5, and contact time of 330 min.
Figure 8. Adsorption isotherms of cheese whey on ACPO4 at different temperatures: (a) lactose; (b) BOD5; (c) COD. Conditions: adsorbent dosage of 2 g L−1, initial pH of 6.5, and contact time of 330 min.
Water 15 03682 g008
Figure 9. Proposed adsorption mechanism for the adsorption of cheese whey onto ACPO4.
Figure 9. Proposed adsorption mechanism for the adsorption of cheese whey onto ACPO4.
Water 15 03682 g009
Figure 10. Recycle test for the adsorption of cheese whey onto ACPO4: (a) lactose; (b) BOD5; (c) COD.
Figure 10. Recycle test for the adsorption of cheese whey onto ACPO4: (a) lactose; (b) BOD5; (c) COD.
Water 15 03682 g010
Table 1. EDS analysis of the adsorbents.
Table 1. EDS analysis of the adsorbents.
AdsorbentComposition (%)
COSiMgKCaP
BSG86.3513.230.420.260.110.74-
Commercial 5 mm86.3513.230.42-
PO454.4730.721.7413.07
Table 2. Textural properties of the adsorbents.
Table 2. Textural properties of the adsorbents.
AdsorbentSo (m2 g−1)Vp (cm3 g−1)dp (nm)
BSG104.30.112.03
Commercial 5 mm377.50.331.10
ACPO4605.10.412.01
Table 3. Results from the literature comparing activated carbons from biomass with H3PO4.
Table 3. Results from the literature comparing activated carbons from biomass with H3PO4.
PrecursorBET Surface Area
(m2 g−1)
Activation Temperature (°C)Reference
BSG605.1400This study
Spent coffee grounds614.8800[62]
Peanut shells590.7400[63]
BSG768.4500[8]
Rice husk residue585.0400[63]
Table 4. Comparison between different techniques described in the literature for cheese whey wastewater treatment.
Table 4. Comparison between different techniques described in the literature for cheese whey wastewater treatment.
TechniqueRemoval (%)Reference
LactoseBOD5COD
Adsorption with activated carbon634691This study
Upflow anaerobic sludge blanket90[82]
Vertically moving biofilm system89[83]
Activated sludge90[23]
Coagulation–flocculation with FeCl3542332[84]
Coagulation–flocculation with Al2(SO4)3493536
Ozone404363[24]
Precipitation with lime564555[85]
Precipitation with NaOH344450
Table 5. Kinetic parameters for the uptake of CW onto ACPO4. Conditions: adsorbent dosage of 2 g L−1, temperature of 25 °C, initial pH of 6.5.
Table 5. Kinetic parameters for the uptake of CW onto ACPO4. Conditions: adsorbent dosage of 2 g L−1, temperature of 25 °C, initial pH of 6.5.
ModelParameterAbsorbate
LactoseBOD5COD
PFOqe3.44 g lactose g−1 adsorbent482.24 mg O2 g−1 adsorbent5234.06 mg O2 g−1 adsorbent
k18.37 × 10−3 min−124.93 × 10−3 min−110.98 × 10−3 min−1
t1/287.10 min28.02 min62.68 min
t0.95359.32 min121.25 min279.52 min
R2adj0.950.950.97
SD0.27 g lactose g−1 adsorbent31.15 mg O2 g−1 adsorbent252.03 mg O2 g−1 adsorbent
BIC−28.5494.54148.90
PSOqe3.52 g lactose g−1 adsorbent518.40 mg O2 g−1 adsorbent5315.40 mg O2 g−1 adsorbent
k21.16 × 10−3 g lactose g−1 adsorbent min−10.06 × 10−3 mg O2 g−1 adsorbent min−10.02 × 10−3 mg O2 g−1 adsorbent min−1
t1/2136.93 min27.09 min61.93 min
t0.95382.95 min308.90 min333.04 min
R2adj0.940.960.98
SD0.30 g lactose g−1 adsorbent18.43 mg O2 g−1 adsorbent171.17 mg O2 g−1 adsorbent
BIC−25.9770.89168.74
AFOqe3.18 g lactose g−1 adsorbent506.44 mg O2 g−1 adsorbent5303.37 mg O2 g−1 adsorbent
kAV19.56 × 10−3 min−110.69 × 10−3 min−15.22 × 10−3 min−1
nAV2.570.470.59
t1/273.88 min24.86 min61.18 min
t0.95200.34 min217.30 min315.99 min
R2adj0.970.970.99
SD0.21 g lactose g−1 adsorbent23.41 mg O2 g−1 adsorbent107.93 mg O2 g−1 adsorbent
BIC−40.3883.13119.81
Table 6. Equilibrium parameters for the uptake of CW onto ACPO4. Conditions: adsorbent dosage of 2 g L−1, initial pH 6.5, and contact time of 330 min.
Table 6. Equilibrium parameters for the uptake of CW onto ACPO4. Conditions: adsorbent dosage of 2 g L−1, initial pH 6.5, and contact time of 330 min.
ModelParameterT (K)
288298308318
Adsorbate
LactoseBOD5CODLactoseBOD5CODLactoseBOD5CODLactoseBOD5COD
LangmuirqmL * 4.55713.507378.635.41988.049050.276.861115.7810,657.127.441349.3112,220.00
kL (L mg−1)0.00010.00040.000020.00050.00040.000040.400.770.110.541.020.16
R2adj0.9950.9990.9950.9980.9980.9810.9960.9970.9730.9980.9930.998
SD0.094.75165.350.068.87432.100.1017.84577.460.7732.43171.29
BIC−20.6018.8054.30−24.8425.0563.90−19.9332.0366.80−22.4038.0054.65
FreundlichkF +1.074.2711.711.886.0529.752.1629.92322.262.8457.06853.09
nF1.461.801.782.291.771.992.132.473.272.472.814.29
R2adj0.9890.9910.9750.9880.9890.9730.9850.9880.9480.9970.9980.978
SD0.1516.48381.150.1527.05511.610.2036.14848.750.1118.59601.47
BIC−15.8331.2462.65−15.6236.1865.56−12.5539.0970.65−19.6332.4467.21
Dubinin–RadushkevichqmDR *4.25496.326219.234.27911.628286.645.171016.979794.238.061726.8911,237.52
β (mol2 kJ−2)31.2325.6631.1920.2717.0020.8021.3212.9211.888.075.3510.59
Es (kJ mol−1)0.130.140.130.160.170.160.150.200.210.250.310.22
R2adj0.9970.9960.9850.9980.9980.9280.9970.9960.9770.9990.9990.999
SD0.0610.50296.430.0610.52833.990.0918.19532.190.0410.35127.37
BIC−24.0722.9749.69−18.0422.9857.97−14.6727.3754.37−21.7922.8542.93
HillqmH *3.69719.896721.234.221043.768229.114.721371.1312,759.1612.773940.9912,857.92
C1/2 (mg L−1)3.172697.2727,607.021.772591.1121,454.731.831549.5911,212.783.454526.726488.04
nH1.360.991.191.161.031.181.250.900.890.750.580.96
R2adj0.9980.9990.9960.9990.9990.9980.9980.9980.9990.9990.9990.999
SD0.075.31146.990.069.75124.250.0918.5795.620.049.63188.63
BIC−27.5017.5144.08−28.6622.3842.74−25.3527.5340.64−28.6922.2846.07
Note(s): * The unit of qmL, qmDF, and qmH for lactose is (g lactose g−1 adsorbent), and for BOD5 and COD, it is (mg O2 g−1 adsorbent). + The unit of kF is ((mg g−1) (mg L−1)−1/n).
Table 7. Thermodynamic parameters for the uptake of CW onto ACPO4.
Table 7. Thermodynamic parameters for the uptake of CW onto ACPO4.
AdsorbateT (K)ΔG° (kJ mol−1)ΔH° (kJ mol−1)ΔS° (J mol−1 K−1)Ea (kJ mol−1)
Lactose288−3.31238.26830.1229.24
298−6.26
308−23.74
318−25.31
BOD5288−5.66237.16829.8721.71
298−5.86
308−25.42
318−26.99
COD2881.51265.72910.3739.25
298−0.15
308−20.44
318−22.09
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Castro, L.E.N.; Matheus, L.R.; Mançano, R.R.; Sganzerla, W.G.; da Rosa, R.G.; Barroso, T.L.C.T.; Ferreira, V.C.; Colpini, L.M.S. Single-Step Modification of Brewer’s Spent Grains Using Phosphoric Acid and Application in Cheese Whey Remediation via Liquid-Phase Adsorption. Water 2023, 15, 3682. https://doi.org/10.3390/w15203682

AMA Style

Castro LEN, Matheus LR, Mançano RR, Sganzerla WG, da Rosa RG, Barroso TLCT, Ferreira VC, Colpini LMS. Single-Step Modification of Brewer’s Spent Grains Using Phosphoric Acid and Application in Cheese Whey Remediation via Liquid-Phase Adsorption. Water. 2023; 15(20):3682. https://doi.org/10.3390/w15203682

Chicago/Turabian Style

Castro, Luiz Eduardo Nochi, Larissa Resende Matheus, Rosana Rabelo Mançano, William Gustavo Sganzerla, Rafael Gabriel da Rosa, Tiago Linhares Cruz Tabosa Barroso, Vanessa Cosme Ferreira, and Leda Maria Saragiotto Colpini. 2023. "Single-Step Modification of Brewer’s Spent Grains Using Phosphoric Acid and Application in Cheese Whey Remediation via Liquid-Phase Adsorption" Water 15, no. 20: 3682. https://doi.org/10.3390/w15203682

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

Article Metrics

Back to TopTop