Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Nov 6;8(46):43674–43689. doi: 10.1021/acsomega.3c05229

Banana Peel Derived Chitosan-Grafted Biocomposite for Recovery of NH4+ and PO43–

Himarati Mondal †,*, Bhaskar Datta †,‡,*
PMCID: PMC10666154  PMID: 38027321

Abstract

graphic file with name ao3c05229_0009.jpg

Biomass-derived adsorbents afford accessible and inexpensive harvesting of nitrogen and phosphorus from wastewater sources. Human urine is widely accepted as a rich source of nitrogen and phosphorus. However, direct use of urine in agriculture is untenable because of its unpleasant smell, pathogen contamination, and pharmaceutical residues. In this work, we have grafted chitosan onto dried and crushed banana peel (DCBP) to generate the biocomposite DCBP/Ch. A combination of FTIR, TGA, XRD, FESEM, EDX, and NMR analyses were used to characterize DCBP/Ch and reveal condensation-aided covalent conjugation between O–H functionalities of DCBP and chitosan. The adsorption performance of DCBP/Ch toward NH4+ and PO43– is in sync with its attractive surface porosity, elevated crystallinity, and thermostability. The maximum adsorption capacity of DCBP/Ch toward NH4+/PO43– was estimated as 42.16/15.91 mg g–1 at an operating pH of 7/4, respectively, and ranks highly when compared to previously reported bioadsorbents. DCBP/Ch performs admirably when tested on artificial urine. While nitrogen and phosphorus harvesting from human urine using single techniques has been reported previously, this is the first report of a single adsorbent for recovery of NH4+ and PO43–. The environmental compatibility, ease of preparation, and economic viability of DCBP/Ch present it as an attractive candidate for deployment in waste channels.

1. Introduction

The unfettered worldwide use of synthetic fertilizers has created immense burden on the energy sector and nonrenewable resources such as phosphate rock and natural gas.15 NH3/NH4+ and PO43– are the primary components of fertilizers that continue to rely on energy-intensive methods of production.1,6 In this context, animal wastes have been envisaged as alternatives to synthetic fertilizers since these are abundant sources of nitrogen and relatively of phosphorus compounds. Generally, an adult healthy human excretes about 550 L of urine annually which is equivalent to 4 kg of nitrogen, 0.4 kg of phosphorus, and 0.9 kg of potassium.7 Despite containing as many as 3000 separate chemical species, the dominant components are water (95%) and urea (2%). Ions such as NH4+, Na+, K+, Ca2+, Mg2+, Cl, SO42–, NO32–, and PO43– are present in variable proportions.8,9 NH4+ and PO43– are also part of waste effluents from industries such as fertilizers, synthetic polymers, and refrigerant gas.10,11 Notably, excessive amounts of NH4+ and PO43– in the environment are understood to pollute the soil, water, and air.1,4,1217 High amounts of aqueous NH4+ and PO43– can cause eutrophication in water bodies, creating challenges for aquatic life.18,19 The recovery of NH4+ and PO43– from domestic and industrial wastes can address environmental safety concerns while offering potential strategies for reducing dependence on synthetic fertilizers. In this context, the direct application of human urine as a fertilizing agent is severely constrained by the presence of pathogens, challenges associated with storage and transport, and social taboos.9,20 WHO advises comprehensive sanitization of human urine with storage for at least six months, before being used as a fertilizer. Several distinctive processes have emerged for recovery of NH4+ and PO43– from human urine that preclude long-term storage. Techniques such as ammonia stripping,21,22 chemical precipitation with NH4MgPO4·6H2O,2325 biological denitrification,26 and electrochemical conversion27,28 offer different routes to access NH4+ and PO43– in urine. These methods vary in the efficiencies of NH4+ and PO43– retrieval and the extent of energy or chemical/material investment necessary. Some of these methods may also ultimately contribute to transfer of pollutants. The safe and commercially viable recovery of NH4+ and PO43– from human urine could be considered as a less explored paradigm of sustainable technologies. In this regard, generic methods that are associated with the separation of ions from specific media could be adapted suitably. Coagulation/flocculation, photocatalysis, ultrafiltration, oxidation, Fenton-biological treatments, nanofiltration, membrane separation, and adsorption are among such methods for the separation of ionic substances.29,30 Not surprisingly, adsorption is an attractive method owing to its simplicity, adaptability, and speed of operation.3134 The separate adsorptive removal of NH4+ and PO43– has been previously reported using low cost natural polymers, agricultural wastes, zeolites, clays, and chemically modified inorganic materials.3546 The simultaneous adsorption of NH4+ and PO43– from urine or industrial effluents is relatively unexplored.

We have recently reported on the use of dried banana peel powder as an effective adsorbent for NH4+.42 While banana peel compositions vary based on the climate and region of cultivation, components with significant presence include pectin (10–21%), cellulose (7.6–9.6%), hemicelluloses (6.4–9.4%), and lignin (6–12%).47 Amino acids such as leucine, valine, phenylalanine, and threonine; dietary fiber; and chlorophyll pigments are other major components.4750 Banana peel is a rich source of phenolic compounds including HCA, F, Fols, and CA.51 Interestingly, almost every part of the banana tree, such as the leaves, pseudostems, stalk, trunks, inflorescence, and piths, is fit for human consumption. The banana peel comprises nearly 35–50% of the total mass of the fruit and is almost universally consigned to waste.47,52 Approximately 36 million tons of banana peels are generated worldwide every year and considering their current fate offer an attractive biomass for suitable chemical manipulation. In this regard, the natural polysaccharide chitosan has been used for calibrating the adsorption behavior of biomass.5355 Chitosan is a prominent naturally occurring polycation, whose charge density is a function of degree of acetylation and pH of solution.56 Therefore, chitosan is pH sensitive, i.e., readily dissolves at acidic pH and is insoluble at alkaline pH.57 The prevalent protonated −NH2 functionalities of chitosan render it soluble in acidic pH and insoluble in alkaline pH.5660 We were interested in combining the pH-responsiveness and stability of chitosan with the adsorptive capabilities of banana peel powder.

In this work, we have developed a pH-responsive chitosan-grafted biocomposite of banana peel powder and have applied it for adsorptive exclusion of both NH4+ and PO43–. To the best of our knowledge, this is the first report of a biocomposite that is suitable for adsorbing both NH4+ and PO43– by variation of the pH of the medium.

2. Experimental Section

2.1. Materials

Banana peels, chitosan (molecular weight = 310 kDa, viscosity = 1000 cP in CH3COOH solution, and degree of deacetylation >75%), CH3COOH, CH3COONa, NH4Cl, Na2SO4, KCl, NaCl, CaCl2, MgSO4·7H2O, NaH2PO4·2H2O, Na2HPO4·2H2O, uric acid (C5H4N4O3), creatinine (C4H7N3O), sodium citrate (Na3C6H5O7·2H2O), potassium oxalate (K2C2O4·H2O), and urea (CO(NH2)2) were purchased from Sigma-Aldrich and used without any modification/purification. All experiments in aqueous media were carried out by using MQW.

2.2. Preparation of DCBP/Ch

For synthesizing DCBP/Ch, first, banana peels were collected from waste stuffs of Indian Institute of Technology-Gandhinagar, washed thoroughly by MQW to remove superficial dirt and sticking impurities, cut into small pieces, sun-dried for 1 week, and finally oven-dried at 60 °C for 2 days. Such dried peels were crushed by using a mortar–pestle to obtain the DCBP powder (Scheme 1), which was used as the primary raw material for synthesizing DCBP/Ch. In the next stage, DCBP/Ch was synthesized by grafting chitosan within the microstructures of DCBP by using the optimum recipe (Table 1). Briefly, 3 mL of glacial CH3COOH was dissolved in 100 mL of MQW and preheated to 40 °C, followed by the gradual addition of 4 g of oven-dried chitosan with constant stirring at 200 rpm. To this viscous chitosan solution, 1.5 g dried of the previously processed DCBP was added with constant stirring at 200 rpm and 25 °C. Stirring was continued for 6 h to ensure maximum grafting of chitosan with DCBP. The as-obtained DCBP/Ch was filtered under suction, washed thoroughly with 3:1 MQW:CH3OH solution, dried in a hot-air oven for 6 h, and then stored in an airtight container for subsequent experiments.

Scheme 1. Synthetic Methodology of DCBP/Ch.

Scheme 1

Table 1. Systematic Optimization of the Synthesis Parameters to Obtain the Optimum DCBP/Ch.

2.2.

#

Not applicable.

2.3. Characterization

Structures of chitosan, DCBP, DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch were characterized by FTIR (Spectrum-2, Singapore, PerkinElmer), TG (TGA 4000, PerkinElmer), XRD (D8 Discover, Bruker), FESEM/EDX (JSM7600F, Jeol), 1H NMR (Ascend NMR-500 MHz, Bruker), DLS (Nano ZS, Malvern Instruments), and UV–vis (Specord 210 Plus, AnalytikJena) analyses. All chemical structures and graphical presentations were generated by ChemDraw Ultra 12.0 and Origin 9.0 software, respectively.

2.4. Adsorption Methodology

Isothermal adsorption experiments were conducted by dipping 0.025 g of dry DCBP/Ch within 50 mL of NH4+/PO43– solution within 5–30/1–10 mg L–1 at the pHo = 7/4. Adsorption isotherm and kinetics data were determined by measuring unadsorbed NH4+/PO43– concentrations (mg L–1) at different time intervals by ammonia electrode (Hanna HI4101) and ICPMS (Nexion 2000B ICP-MS, PerkinElmer). The adsorption capacities (ACs, mg g–1), rate constants, and thermodynamics parameters were estimated by methods reported elsewhere.61,62 Briefly, adsorption capacity at time t (qt, mg g–1) and equilibrium (qe, mg g–1) were calculated by using eqs 1 and 2, respectively.

2.4. 1
2.4. 2

Here, C0, Ct/Ce, V, and ms represent the initial concentration, residual concentrations at time t/equilibrium, volume of adsorbate solution, and mass of adsorbent, respectively. Separate calculations were performed from each replicate of adsorption experiments.

2.5. Preparation of ArUr

The principal components of human urine are provided in Table S1. In 100 mL of MQW, all the components listed in Table S1 were dissolved in isothermal conditions (37 ± 0.2 °C) by using a magnetic stirrer-cum-hot plate at 250 rpm. The final solution was stored in a tightly sealed flask. The ArUr thus prepared was used within 24 h. The ingredients selected for preparation of the artificial urine include only those substances that are present at relatively high concentrations that do not vary significantly in human urine.42

3. Results and Discussion

3.1. Characterization to Confirm the Formation of DCBP/Ch from DCBP and Chitosan

3.1.1. FTIR Analyses of Chitosan, DCBP, and DCBP/Ch

FTIR characterization was performed on DCBP and is displayed in Figure 1a. The band at 1740 cm–1 is attributed to the C=O str. of −COOH/–COOCH3.63 The presence of cellulose, hemicellulose, and lignin in DCBP can be inferred from the bands at 2921/2852 and 1368–1145 cm–1 attributable to C–H str. and bending, respectively.33,63,64 In our previous work we have described the FTIR spectra of BNPP formulation.42 Modifications in the process of drying and crushing deployed in the current work likely impose subtle variations in the behavior of chemical functionalities in FTIR. For instance, the greater population of O–H functionalities in DCBP compared to BNPP could be understood from the shifting of broad mutually hydrogen-bonded O–H···N–H str. from 3286 cm–1 in BNPP42 to 3315 cm–1 in DCBP6469 as well as the appearance of O–H str. of amino acids at 2970 cm–1 in DCBP.63 Notwithstanding these subtle variations in functional group composition of DCBP, the intimate association of chitosan with DCBP in DCBP/Ch could be readily inferred from multiple FTIR bands. Notably, sharpening and shifting of mutual hydrogen-bonded O–H···N–H str. was observed from 3361 and 3315 cm–1 of chitosan and DCBP, respectively, to 3449 cm–1 in DCBP/Ch.6469 The O–H str. of amino acids at 2970 cm–1 in DCBP remain unaltered in DCBP/Ch, and the nonavailability of this peak in chitosan indicates the presence of a DCBP moiety in DCBP/Ch. The insertion of DCBP in DCBP/Ch was also understood from C–H str. of alkane and C=O str. of −COOH/–COOCH3 at 2921/2852 and 1740 cm–1 in DCBP and 2925 and 1739 cm–1 in DCBP/Ch, respectively, and the absence of the corresponding peaks in chitosan.33,63,64 Conversely, CH2str. of the pyranose ring of chitosan at 2866 cm–1 could be observed at 2865 cm–1 in DCBP/Ch and substantiate the presence of chitosan in DCBP/Ch.64 The O–C coupled association of chitosan with DCBP in DCBP/Ch is further confirmed from the emergence of a peak at 1119 cm–1 in DCBP/Ch.70 This band was attributed to the newly generated >CH–O–CH2– linkage in DCBP/Ch. The disappearance of C–OH def. of chitosan at 1420 cm–1 in DCBP/Ch indicated the conversion of C–OH into C–O–C moieties.64 The FTIR analysis of DCBP and DCBP/Ch is detailed as part of Table 2.

Figure 1.

Figure 1

FTIR analyses of (a) chitosan, DCBP, and DCBP/Ch and (b) DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch.

Table 2. FTIR Analyses.
Chitosan DCBP DCBP/Ch N-DCBP/Ch P-DCBP/Ch Assignment
3361 3315 3449 3297 3314 mutual OH···NH hydrogen bonding in chitosan and O–H group of polymeric compounds in DCBP6469
2970 2970 O–H str. of amino acid groups66
2921, 2852 2925 2924 2922 C–H asym. str. of >CH–33,64
2866 2865 2862 2860 CH2str. of pyranose ring64
1740 1739 1738 1738 C=O str. of −COOH/–COOCH3 indicating the presence of carboxylic acids and/or their esters63
1625 1624 N–H def. in amine63
1621, 1530 C=O asym. str. of –COO70,87
1588 1594 1596 1589 N–H bending(63)
1460 1462 1460 aromatic ring vibration of lignin63
1453 C=O sym. str. of –COO87
1420   C–OH def.(64)
1380 1374 1375 1375 C–H twisting of >CH264
1368–1145 1364–1147 1314–1146 1315–1142 C–H bending of crystalline cellulose and C–H bending of cellulose, hemicelluloses, and lignin polymer63
1263 –CH2OH (side chain) of chitosan64
1154 1155 1152 asym. str. of C–O–C and C–N str.(67)
1119 1121 1123 >CH–O–CH2–, originated from grafting70
1046 1047 1047 1047 C–N str. of DCBP68
1065, 1026 1030 1066–1037 1068–1035 skeletal vibration of C–O str.(67)
988 P–O–H bond70,88,89
809–567 889, 780, 719 885–765 C–H vibration of carbohydrates69
713–560 549, 521 519 skeletal mode of pyranose ring64

3.1.2. TG Analyses of Chitosan, DCBP, and DCBP/Ch

Thermogravimetry is an important tool for the study of biopolymeric composites via a scrutiny of changes in thermal stabilities. We performed TG analyses of chitosan, DCBP, and DCBP/Ch to observe the relative change in thermal stability of DCBP/Ch because of grafting. The thermograms of chitosan, DCBP, and DCBP/Ch (Figure S1a) indicate three major decomposition ranges. Of these, the first range between 30 and 105 °C corresponds to elimination of volatile water.64 This faster degradation rate for DCBP in this stage corresponds to a lower extent of hydrogen bonding within hydrophilic functionalities, such as −COOH and O–H. The second range in the thermograms between 238 and 386 °C represents the loss of H2O during anhydride formation. The higher population of −COOH groups in DCBP accompanied by lower hydrogen bonding likely leads to rapid degradation in this region. Notably, the thermal stability of DCBP/Ch is significantly higher compared to DCBP, confirming the association of chitosan with DCBP. Moreover, the slower skeletal decomposition and higher residue in DCBP/Ch (24.57 wt %) than that of DCBP (21.72 wt %) indicates an improvement in structural stability of DCBP/Ch because of grafting-related covalency and hydrogen bonding.71

3.1.3. XRD Analyses of Chitosan, DCBP, and DCBP/Ch

The grafting of chitosan with DCBP is expected to alter Miller planes of DCBP. Observing the changes in the crystallinity of DCBP after chitosan grafting can indicate association between the two components. The sharp peak at 2θ = 10.15° (Figure 2a) observed for DCBP is attributed to the hydrogen-bonding-aided orderly arrangement of Miller indices. In contrast, two major peaks were observed at 2θ = 11.08° and 20.65°, for DCBP/Ch. The peak at 2θ = 11.08° is attributed to hydrogen-bonded orderly arrangement of planes from DCBP and chitosan. The peak at 2θ = 20.38° originates from the polymer-specific peak of chitosan. The shifting of the weak shoulder at 2θ = 38.01° of chitosan in XRD of DCBP/Ch indicates association of chitosan with DCBP.

Figure 2.

Figure 2

XRD analyses of (a) DCBP, chitosan, and DCBP/Ch and (b) DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch.

3.1.4. FESEM and EDX Analyses of Chitosan, DCBP, and DCBP/Ch

The grafting of chitosan with DCBP could affect the surface morphology of the latter. FESEM images of DCBP suggest a porous and fibrous surface morphology (Figure 3a). In contrast, chitosan displayed a mostly dense and homogeneous surface (Figure 3b). Notably, the fibrous/porous and dense/homogeneous characteristics of DCBP and chitosan, respectively, coexist in DCBP/Ch (Figure 3c,d). These results are consistent with the FTIR analysis and together with the XRD and TGA results justify intimate association of chitosan with DCBP in DCBP/Ch. From the EDX plot of DCBP/Ch (inset of Figure 3c), distinct peaks of C, N, and O were observed.

Figure 3.

Figure 3

FESEM images of (a) DCBP, (b) chitosan, (c and d) DCBP/Ch, (e) N-DCBP/Ch, and (f) P-DCBP/Ch and EDX of (c) DCBP/Ch, (e) N-DCBP/Ch, and (f) P-DCBP/Ch in the insets.

3.1.5. NMR Analyses of Chitosan, DCBP, and DCBP/Ch

NMR has been used for elucidating structural nuances of natural polymers, biomass, and chemically modified biomass.7277 Based on previous reports on the composition of the banana peel, the DCBP is likely to be comprised of pectin, cellulose, hemicellulose, lignin, amino acids, and phenolic compounds. Of these, pectin contains α-d-GalA, β-d-Galp, α-l-Araf, and α-l-Rhap. The 1H NMR spectrum of DCBP is shown in Figure 4 and described in Table 3. Close scrutiny of the 1H NMR of DCBP revealed the presence of various constituent sugars of pectin. Notably, H1, H2, H3/H4, and H5 of α-d-GalA; H1 and H2/H3/H4 of α-l-Araf; H1 and H2/H3/H5 of β-d-Galp; and −CH3 in α-l-Rhap were observed at 4.45, 3.12, 3.24–3.61, and 3.89; 5.72 and 3.89; 4.52 and 3.24–3.61; and 1.24 ppm, respectively.32,78,79 Similarly, the presence of cellulose in DCBP could be inferred from the characteristic peaks of H1, H2, H3/H4/H5, and H6 at 4.10, 2.65, 3.07, and 3.61–3.89 ppm, respectively.80 Hemicellulose contains β-d-xylp, d-GlcA, and α-l-Araf. The 1H NMR spectrum of DCBP displayed characteristic hemicellulose-specific chemical shifts, such as H1, H2/H3/H4, and H5 of β-d-Xylp; H1 and O–CH3 of d-GlcA; and H1 and H2/H3/H4 of α-l-Araf at 4.45, 3.24–3.61, and 4.03; 5.35 and 1.35; and 5.72 and 3.89 ppm, respectively (Table 3).8183 The presence of lignin in DCBP was confirmed from peaks of oxygenated aliphatic, aromatic, and O–CH3 protons at 2.18/2.38/3.61–3.89/5.35, 5.52/6.63/6.72/6.91 and 3.89 ppm, respectively.72,84 The presence of Leu, Val, Phe, and Thr could be inferred from characteristic chemical shifts of each amino acid.48 These include α-H, β-H, γ-H, and δ-H of Leu; β-H and γ-H of Val; α-H, β-H, and phenyl ring-H of Phe; α-H, β-H, β-OH, and γ-H of Thr; and N-H of Leu/Val/Phe/Thr observed at 3.61, 1.82, 1.49, and 0.91; 2.36 and 0.91; 4.10, 3.07, and 6.63–6.91; 3.61, 3.89, 3.61, and 1.24; and 6.02 ppm, respectively. Further, the incidence of phenolic compounds such as HCA, F, Fols, and CA was established by characteristic chemical shifts in the 1H NMR of DCBP.51,85 These include phenyl ring-H of HCA/F/Fols/CA and vinylic-H of HCA; O–H of the aliphatic and aromatic region in Fols/CA and HCA/CA; −CHa2–CHb(OH)– of Fols; and NHc2–CHd2–CHe(OH)– of CA observed at 6.02–6.91; 3.61 and 5.35; 2.65 (a) and 4.45 (b); 6.02 (c), 3.07 (d), and 4.71 (e) ppm, respectively.

Figure 4.

Figure 4

1H NMR of DCBP.

Table 3. 1H NMR Analyses.
Chemical Shift (ppm)
 
Chitosan DCBP DCBP/Ch Explanation
4.61 4.63 H1 of chitosan64,86
2.43 H2 of chitosan64
3.67 3.65 H3 of chitosan64,86
3.82 3.85 H4 of chitosan64,86
3.54 H5 of chitosan64,86
3.91 3.95 H6 of chitosan64,86
4.92 4.91 –NH2 of chitosan64
3.25 –CH2OH of chitosan64,86
4.45 H1 of α-d-GalA78,79
3.12 3.17 H2 of α-d-GalA78,79
3.24–3.61 3.22–3.65 H3/H4 of α-d-GalA32,78,79
3.89 3.85 H5 of α-d-GalA78,79
5.72 5.64 H1 of α-l-Araf78,79
3.89 3.85 H2/H3/H4 of α-α-l-Araf78,79
4.52 H1 of β-d-Galp78,79
3.24–3.61 3.22–3.65 H2/H3/H5 of β-d-Galp78,79
1.24 1.23 –CH3 of α-l-Rhap78,79
4.10 4.12 H1 of cellulose80
2.65 2.61 H2 of cellulose80
3.07 H3/H4/H5 of cellulose80
3.61–3.89 3.65–3.85 H6 of cellulose80
4.45 4.46 H1 of β-d-xylp in hemicellulose81
3.24–3.61 3.22–3.65 H2/H3/H4 of β-d-xylp in hemicellulose81
4.03 H5 of β-d-xylp in hemicellulose81
5.35 5.31 H1 of d-GlcA in hemicellulose81,82
1.35 1.34 O–CH3 of d-GlcA in hemicellulose82
5.72 5.64 H1 of α-l-Araf in hemicellulose83
3.89 3.85 H2/H3/H4 of α-l-Araf in hemicellulose83
2.18/2.38/3.61–3.89/5.35 2.14/2.39/3.65–3.85/5.31 oxygenated aliphatic region in lignin72,84
5.52/6.63/6.72/6.91 5.64/6.70/6.88/6.97 aromatic region in lignin72,84
3.89 3.85 O–CH3 of lignin
3.61 3.65 α-H of leu
1.82 1.85 β-H of leu
1.49 1.48 γ-H of leu
0.91 0.94 δ-H of leu
2.36 2.32 β-H of val
0.91 0.94 γ-H of val
4.10 4.12 α-H of phe
3.07 β-H of phe
6.63–6.91 6.70–6.97 phenyl ring-H of phe
3.61 3.65 α-H of thr
3.89 3.85 β-H of thr
3.61 3.65 β-OH of thr
1.24 1.23 γ-H of thr
6.02 N-H of leu/val/phe/thr
6.02–6.91 6.70–6.97 phenyl ring-H of HCA/F/Fols/CA and vinylic-H of HCA
3.61 3.65 OH peaks of aliphatic region in Fols/CA
5.35 5.31 OH peaks of aromatic region in HCA/CA
2.65 2.61 –CH2–CH(OH)– of Fols
4.45 4.46 –CH2–CH(OH)– of Fols
6.02 NH2–CH2–CH(OH)–
3.07 NH2–CH2–CH(OH)–
4.71 4.70 NH2–CH2–CH(OH)–

The 1H NMR of DCBP/Ch is expected to indicate the presence of both primary components, namely, the DCBP and chitosan. Scrutiny of the 1H NMR of chitosan led to the attribution of H1, H2, H3, H4, H5, H6, −NH2, and −CH2OH peaks of chitosan to the observed chemical shifts at 4.61, 2.43, 3.67, 3.82, 3.54, 3.91, 4.92, and 3.25 ppm (Figure S2 and Table 3), respectively.64,86 In the case of our chitosan-grafted DCBP, 1H NMR of the composite showed peaks at 4.63, 3.65, 3.85, 3.95, and 4.91 ppm (Figure 5 and Table 3). Based on comparison with cognate peaks of chitosan, the above peaks could be attributed to H1, H3, H4, H6, and −NH2, respectively, and also indicate the successful incorporation of chitosan in DCBP/Ch. The presence of pectin-, cellulose-, hemicellulose-, lignin-, amino acid-, and phenolic compound-specific 1H NMR peaks was similarly confirmed in DCBP/Ch. For instance, H2, H3/H4, and H5 of α-d-GalA; −CH3 of α-l-Rhap; H1 and H2/H3/H4 of α-l-Araf; and H2/H3/H5 of β-d-Galp peaks were observed at 3.17, 3.22–3.65, and 3.85; 1.23; 5.64 and 3.85; and 3.22–3.65 ppm, respectively, and indicated the presence of pectin in DCBP/Ch. Further, H1/H2/H6 of cellulose and oxygenated aliphatic, aromatic, and O–CH3 protons of lignin were inferred from peaks at 4.12/2.61/3.65–3.85 and 2.14/2.39/3.65–3.85/5.31, 5.64/6.70/6.88/6.97, and 3.85 ppm, respectively, thereby confirming the presence of cellulose and lignin in DCBP/Ch. Likewise, the prevalence of hemicellulose in DCBP/Ch was affirmed from H1 and H2/H3/H4 of β-d-Xylp; H1 and O–CH3 of d-GlcA; and H1 and H2/H3/H4 of α-l-Araf peaks at 4.46 and 3.22–3.65; 5.31 and 1.34; and 5.64 and 3.85 ppm, respectively. The amino acid residual peaks, such as α-H/β-H/γ-H/δ-H of Leu, β-H/γ-H of Val, α-H/phenyl ring-H of Phe, and α-H/β-H/β-OH/γ-H of Thr peaks at 3.65/1.85/1.48/0.94, 2.32/0.94, 4.12/6.70–6.97, and 3.65/3.85/3.65/1.23 ppm, respectively, corroborated the presence of amino acids of DCBP in DCBP/Ch. Also, the presence of phenolic compounds of DCBP in DCBP/Ch was supported by peaks at 6.70–6.97; 3.65 and 5.31; 2.61 (a) and 4.46 (b); and 4.70 (c) ppm attributable to the phenyl ring-H of HCA/F/Fols/CA and vinylic-H of HCA; O–H of the aliphatic and aromatic region in Fols/CA and HCA/CA; −CHa2–CHb(OH)– of Fols; and −CHc(OH)– of CA, respectively.

Figure 5.

Figure 5

1H NMR of DCBP/Ch.

The grafting of chitosan with DCBP through condensation-type association between −CH2OH of chitosan and >CHOH of α-d-GalA of the pectin moiety of DCBP was inferred from the newly generated >CHa–O–CHb2– linkages in DCBP/Ch at 3.17(a) and 3.95(b) ppm, respectively. The chemical shifts of these protons are consistent with the previously described FTIR peak of >CH–O–CH2– at 1119 cm–1 in DCBP/Ch and the disappearance of C–OH def. of chitosan at 1420 cm–1 in DCBP/Ch. The grafting-aided modification in thermal properties, crystallinity, and surface morphologies were also confirmed from TG, XRD, and FESEM analyses. Moreover, the increase in hydrodynamic radius of DCBP/Ch compared to DCBP (Figure S5a–d) and distinct changes in λmax values in the corresponding UV–vis spectra also support the bonding of chitosan with DCBP (Figure S5e).

3.2. Characterization to Confirm the Mechanism of Adsorption of NH4+ and PO43– by DCBP/Ch

3.2.1. FTIR Analyses of N-DCBP/Ch and P-DCBP/Ch

The FTIR spectrum of DCBP/Ch (Figure 1b) after NH4+ adsorption at pHo > pHPZC suggested participation of specific groups in the chitosan-grafted polymer. The hydrogen-bonded mutual O–H···N–H str. at 3449 cm–1 became very sharp in N-DCBP/Ch and shifted to 3297 cm–1. Such a strong bathochromic shift confirmed the increase in the population of N–H functionalities after NH4+ adsorption. Interestingly, the sharp C=O str. of −COOH in DCBP/Ch became very weak in N-DCBP/Ch. This is expected since adsorption was carried out at pHo > pHPZC, where most of the constituent −COOH is deprotonated to −COO. The higher population of −COO could also be inferred from the appearance of new peaks at 1621/1530 and 1453 cm–1 in N-DCBP/Ch,70,87 attributed to νas(−COO) and νs(−COO), respectively. The association of −COO with NH4+ is likely justified by a combination of I and BC modes. In addition, the observed C–N str. at 1047 cm–1 for DCBP/Ch became very sharp in N-DCBP/Ch, confirming the greater accumulation of amine functionalities.

In contrast to the adsorption of NH4+, the adsorption of PO43– by DCBP/Ch was carried out at a diametrically opposite pHo, i.e., pHo < pHPZC. Hydrogen bonds were purported to become stronger at this pHo, and free −NH2 functionalities of chitosan would be protonated to generate NH4+ ions, providing for ideal adsorption conditions for PO43–. The sharpening of hydrogen-bonded mutual O–H···N–H str. at 3449 cm–1 was observed for DCBP/Ch upon adsorption of PO43–. However, in contrast to N-DCBP/Ch, this hydrogen-bonded mutual O–H···N–H str. exhibited hypsochromic shifting to the center at 3314 cm–1, thereby confirming the decrease in relative population of N–H functionalities upon binding with PO43–. Also, the appearance of a new sharp peak at 988 cm–1 indicated the formation of a P–O–H bond in P-DCBP/Ch.70,88,89

3.2.2. TG Analyses of N-DCBP/Ch and P-DCBP/Ch

In N-DCBP/Ch, the conversion of most of −COOH/–OH into −COO/–O at pHo > pHPZC would likely result in more rapid elimination of water due to the reduced extent of hydrogen bonding compared to DCBP/Ch. However, since PO43– adsorption was carried out at pHo < pHPZC, the extent of hydrogen bonding was expected to be retained in P-DCBP/Ch. In the next stage of degradation (243–388 °C) (Figure S1b), elimination of H2O through anhydride formation was restricted in N-DCBP/Ch because of the participation of −COO in ionic/coordinate bonding with NH4+. In contrast, since −COOH is unlikely to be involved with adsorption of PO43–, no gain in thermostability was observed in this region for P-DCBP/Ch. However, the much slower skeletal decomposition of P-DCBP/Ch than those of DCBP/Ch and N-DCBP/Ch and the higher residue in P-DCBP/Ch (i.e., 35.22 wt %) compared to DCBP/Ch and N-DCBP/Ch (24.57 and 18.49 wt %, respectively) suggested the stronger bonding of PO43– with DCBP/Ch than that of NH4+ along with the prevalence of the corresponding metal ion, i.e., Na. In N-DCBP/Ch, skeletal backbone decomposition was maximum because of the evaporation of NH3. Therefore, from thermal analysis, simultaneous adsorption of NH4+ and PO43– could be envisaged.

3.2.3. XRD Analyses of N-DCBP/Ch and P-DCBP/Ch

XRD of N-DCBP/Ch was signified by the complete disappearance of the hydrogen-bonding-specific XRD peak at 2θ = 11.08° (Figure 2b). This is justified by the weakening of hydrogen bonds because of the deprotonation of −COOH/–OH functionalities in N-DCBP/Ch at pHo > pHPZC. Notably, the polymer-specific peak of DCBP/Ch at 2θ = 20.65° was retained at 2θ = 21.67° in N-DCBP/Ch.

The XRD spectrum of P-DCBP/Ch showed major peaks at 2θ = 10.76, 22.58, 32.21, and 33.24°. Since, adsorption of PO43– was carried out at pHo < pHPZC, better orderly arrangement of Miller indices was achieved in P-DCBP/Ch, which resulted in the strengthening of hydrogen bonding. The peak at 2θ = 10.76° indicated hydrogen-bonding interactions within hydrophilic functionalities. The peak at 2θ = 22.58° was the characteristic polymer peak, and the last two peaks appeared because of the Na+ counterions coming from Na2HPO4.

3.2.4. FESEM and EDX Analyses of N-DCBP/Ch and P-DCBP/Ch

Distinct changes in the surface architecture of DCBP/Ch after NH4+ and PO43– adsorption could be visualized via FESEM (Figure 3e,f). Though the heterogeneity of DCBP/Ch persisted after NH4+ and PO43– adsorption, the entire surface was found to be covered by superficial depositions of NH4+ and PO43–. In fact, the fibrous structure of DCBP/Ch could not be visualized upon NH4+ and PO43– adsorption, suggesting thick coverage of adsorbates. This observation was further justified from the higher intensity of the N-specific peak and arrival of the new P-specific peak in EDX plots of N-DCBP/Ch and P-DCBP/Ch (inset of Figure 3e,f), respectively.

In this context, the retention of chitosan after NH4+ and PO43– adsorption could be inferred via chitosan-specific FTIR peaks, such as −CH2str. of the pyranose ring at 2862 and 2860 cm–1 in N-DCBP/Ch and P-DCBP/Ch, respectively, as well as grafting-specific −CH2str. of >CH–O–CH2– at 1121 and 1123 cm–1 in N-DCBP/Ch and P-DCBP/Ch, respectively. Moreover, the polymer-specific peak of DCBP/Ch at 2θ = 20.65° was retained at 2θ = 21.67 and 22.58° in N-DCBP/Ch and P-DCBP/Ch, respectively, suggesting the unaltered polymer structure, i.e., the integrity of chitosan postadsorption. The presence of chitosan after adsorption was evident from the improved thermostability beyond 300 °C compared with DCBP, very much similar to DCBP/Ch. The elevated hydrodynamic radii (Rh) of DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch, than that of DCBP, also indicated the presence of chitosan after adsorption. The persistence of heterogeneity of DCBP/Ch in N-DCBP/Ch and P-DCBP/Ch indicated no change in the composition of the complex after NH4+ and PO43– adsorption. Finally, the observation of chitosan-specific 1H NMR peaks at 3.66, 3.81, 3.96, 4.64, and 4.93 ppm (Figure S3) and 3.17, 3.66, 3.83, 3.96, 4.64, and 4.91 ppm (Figure S4) in N-DCBP/Ch and P-DCBP/Ch, respectively, precludes the washing out of chitosan during adsorption.

3.3. Adsorption of NH4+ and PO43– by DCBP/Ch

The equilibrium adsorption data were fitted to linear and nonlinear forms of different isotherm models such as Langmuir, Henry, Freundlich, Sips, and BET (eqs 37).42,78 The Langmuir model provided the best fits for both NH4+ and PO43– adsorption, based on the highest adjusted R2/F and lowest χ2 values.90

3.3. 3
3.3. 4
3.3. 5
3.3. 6
3.3. 7

Here, qe, Ce, and n/qBET/γ represent adsorption capacity, residual NH4+ and PO43– concentration at equilibrium, and Freundlich/BET/Sips parameters, respectively. kF, k1/k2, kS, kH, and kL are Freundlich, BET, Sips, Henry, and Langmuir isotherm constants, respectively.

Herein, the adsorption of NH4+ and PO43– has been carried out at pHo = 7 and 4, which is higher and lower than that of pHPZC of DCBP/Ch (Figure S6). The maximum AC values, i.e., qmax, of NH4+ adsorption were estimated to be 41.97/42.16, 38.45/39.11, 36.02/36.74, and 34.73/34.65 mg g–1 from linear/nonlinear fitting at 298, 303, 308, and 313 K, respectively (Figures 6a and S7a and Table 4). Similarly, the qmax values of PO43– adsorption were estimated to be 15.91/12.43, 15.52/10.59, 15.09/10.09, and 14.29/9.95 mg g–1 from nonlinear/linear fitting at 298, 303, 308, and 313 K, respectively (Figures 6b and S7c and Table 4). In this context, the Langmuir isotherm model was reported as the best fit model during adsorption of NH4+ by various adsorbents, such as chitosan- and zeolite-based compounds, natural turkish (Yıldızeli) zeolite, natural chinese zeolite, transcarpathian clinoptilolite, natural turkish zeolite, and fruit peels.3541 Similarly, the Langmuir isotherm model was utilized for explaining the adsorption of PO43– by adsorbents, such as modified and virgin tangerine peel biochars, magnetite-incorporated orange peel biochar, and pristine and chemically modified bentonites.4346 The superior modeling of observed kinetics with a pseudo-second-order kinetics model compared to pseudo-first-order kinetics models in both linear/nonlinear forms points to the chemisorption of NH4+ (Figures 6d and S7b) and PO43– (Figures 6c and S7d) by DCBP/Ch.91 The activation energies for NH4+ and PO43– adsorption were estimated to be 70.03 and 35.62 kJ mol–1, respectively (Figure S8a).91 Based on the pore-diffusion model of Weber and Morris, three stages of diffusion patterns were inferred for both NH4+ and PO43– adsorption by DCBP/Ch, suggesting the coexistence of bulk, film, as well as pore/intraparticle diffusions (Figure S9).92 From the slope of the second phase of diffusion, rate constants of intraparticle diffusion, i.e., kip, were calculated. The kip values were found to be 0.7207 and 0.4927 mg g–1 min–0.5, respectively. The lower kip of PO43– indicated the slower intraparticle diffusion rate than that of NH4+, likely due to the larger ionic radius of PO43–. The k2 values were found to increase with the increase in temperature, confirming faster adsorption at higher temperatures. Transformations of nonlinear Langmuir isotherm and pseudo-second-order kinetics equations to linear forms not only implicitly adjust their error structures but also violate error variance. Therefore, recent investigations have suggested the better applicability of nonlinear fitting over linear fitting for kinetics and isotherm data.93 Thus, we have considered all adsorption isotherm and kinetics parameters of nonlinear fitting. The adsorption thermodynamic parameters for N-DCBP/Ch and P-DCBP/Ch are listed in Table 5. The negative values of ΔH° and ΔG° throughout the entire working temperature range clearly indicate the exothermic and spontaneous character of chemisorption underlying N-DCBP/Ch and P-DCBP/Ch (Figures 6e and S8b).94 We have used NH4+ concentrations in this work that are comparable to NH4+ in human urine, as also in our previous work.42 As mentioned in Table S1, the total amount of NH4+ was 1.651 g/100 mL. The amount of PO43– used in our current work (Table S1, 0.374 g/100 mL) is higher than the PO43– levels in human urine. This is because our intent in the current work was to simulate a substantive difference in concentration between NH4+ and PO43–, as present in human urine, and develop an adsorbent that is capable of harvesting both. We are currently working on refining the adsorbent material that could work effectively with the precise concentration of PO43–.

Figure 6.

Figure 6

(a/d/e) Nonlinear isotherm, nonlinear kinetics, and thermodynamics fitting of NH4+ adsorption and (b/c) nonlinear isotherm and nonlinear kinetics fitting of PO43– adsorption.

Table 4. Adsorption Isotherm and Kinetics Parameters.

Temperature (K)
Model/Parameters 298 303 308 313
N-DCBP/Ch
Langmuir (Linear)
qmax (mg g–1)/pHi/C0 (mg L–1) 41.97/7/5–30 38.45/7/5–30 36.02/7/5–30 34.73/7/5–30
kL (L mg–1) 1.2876 1.0102 0.8118 0.7603
R2/F 0.9962/3116.83 0.9994/19722.23 0.9965/3647.61 0.9948/2485.61
χ2 0.8365 0.1142 0.5431 0.7437
Langmuir (Nonlinear)
qmax (mg g–1)/pHi/C0 (mg L–1) 42.16/7/5–30 39.11/7/5–30 36.74/7/5–30 34.65/7/5–30
kL (L mg–1) 1.2412 0.9419 0.7551 0.5841
R2/F 0.9975/3710.43 0.9988/8701.93 0.9953/2280.83 0.9914/1259.46
χ2 2.35 × 10–6 1.07 × 10–6 4.37 × 10–6 8.28 × 10–6
Pseudo-second-order (Linear)
qe,cal (mg g–1)/pHi/C0 (mg L–1) 40.01/7/30 36.41/7/30 32.89/7/30 31.51/7/30
qe,exp (mg g–1) 38.51 ± 0.39 35.62 ± 0.35 33.63 ± 0.38 31.10 ± 0.25
k2 (g mg–1 min–1) 0.0016 0.0026 0.0044 0.0056
R2/F 0.9883/5501.01 0.9841/4301.59 0.9789/3258.39 0.9988/61389.96
χ2 1.4849 1.6685 1.8592 0.0937
Pseudo-second-order (Nonlinear)
qe,cal (mg g–1)/pHi/C0 (mg L–1) 39.04/7/30 35.97/7/30 33.56/7/30 31.25/7/30
qe,exp (mg g–1) 38.51 ± 0.39 35.62 ± 0.35 33.63 ± 0.38 31.10 ± 0.25
k2 (g mg–1 min–1) 0.0022 0.0031 0.0030 0.0071
R2/F 0.9998/73026.96 0.9999/182160.61 0.9999/122195.78 0.9999/1290000
χ2 0.0138 0.0065 0.0111 0.0012
P-DCBP/Ch
Langmuir (Linear)
qmax (mg g–1)/pHi/C0 (mg L–1) 12.43/4/1–10 10.59/4/1–10 10.09/4/1–10 9.95/4/1–10
kL (L mg–1) 1.3526 1.8697 2.1592 2.3455
R2/F 0.9903/600.55 0.9789/284.01 0.9757/247.62 0.9804/311.10
χ2 0.0003 0.0008 0.0009 0.0006
Langmuir (Nonlinear)
qmax (mg g–1)/pHi/C0 (mg L–1) 15.91/4/1–10 15.52/4/1–10 15.09/4/1–10 14.29/4/1–10
kL (L mg–1) 0.8597 0.8522 0.9019 1.0272
R2/F 0.9986/9572.18 0.9956/2105.50 0.9944/1663.51 0.9914/1089.85
χ2 0.0309 0.0934 0.1166 0.1745
Pseudo-second-order (Linear)
qe,cal (mg g–1)/pHi/C0 (mg L–1) 12.51/4/10 12.36/4/10 12.09/4/10 12.06/4/10
qe,exp (mg g–1) 12.37 ± 0.23 12.23 ± 0.25 12.03 ± 0.29 11.95 ± 0.21
k2 (g mg–1 min–1) 0.0092 0.0105 0.0222 0.0102
R2/F 0.9998/63302.42 0.9999/113482.23 0.9999/180018.11 0.9999/88714.98
χ2 0.1542 0.0879 0.0572 0.1179
Pseudo-second-order (Nonlinear)
qe,cal (mg g–1)/pHi/C0 (mg L–1) 12.86/4/10 12.52/4/10 12.41/4/10 11.90/4/10
qe,exp (mg g–1) 12.37 ± 0.23 12.23 ± 0.25 12.03 ± 0.29 11.95 ± 0.21
k2 (g mg–1 min–1) 0.0056 0.0074 0.0091 0.0111
R2/F 0.9943/12251.15 0.9930/10352.97 0.9783/3604.18 0.9741/2573.09
χ2 0.0708 0.0817 0.2451 0.3041

Table 5. Adsorption Thermodynamics Parameters for N-DCBP/Ch and P-DCBP/Ch.

Concentration (ppm)/temperature (K) ΔG° (kJ mol–1) of N-/P-DCBP/Ch –ΔH° (kJ mol–1) of N-/P-DCBP/Ch –ΔS° (J mol–1 K–1) of N-/P-DCBP/Ch
5/298 9.03/– 47.07/– 127.37/–
5/303 8.39/–
5/308 7.88/–
5/313 7.51/–
10/298 8.02/2.87 36.55/4.72 95.34/6.05
10/303 7.66/2.89
10/308 7.25/2.88
10/313 6.97/2.93
15/298 7.34/– 45.39/– 127.96/–
15/303 6.45/–
15/308 5.72/–
15/313 5.81/–
20/298 6.06/– 43.88/– 127.72/–
20/303 4.86/–
20/308 4.17/–
20/313 4.46/–
25/298 4.26/– 32.05/– 93.22/–
25/303 3.72/–
25/308 3.37/–
25/313 3.01/–
30/298 3.11/– 26.27/– 77.64/–
30/303 2.70/–
30/308 2.37/–
30/313 2.06/–

3.4. Binary Adsorption of NH4+ and PO43– from Artificial Urine by DCBP/Ch

Urine is a highly nutrient-dense body fluid containing more than 3000 metabolites.95 The direct use of urine in research on nitrogen and phosphorus harvesting is constrained by the risk of pathogenic activities.20 ArUr has been used as a replacement in academic research. We used ArUr to demonstrate the potential of DCBP/Ch in adsorptive recovery of NH4+ and PO43– simultaneously, at pHo > pHPZC and pHo < pHPZC, respectively.96 Similar to our previous work,42 three different dilutions of ArUr were used for these experiments, namely, 1266 ppm (undiluted/UArUr), 105.48 ppm (1/12 diluted/112ArUr), and 52.74 ppm (1/24 diluted/124ArUr). Also, similar to the adsorption conditions used for adsorbing NH4+ and PO43– from NH4Cl and Na2HPO4, respectively, 25 mg of DCBP/Ch was immersed in 50 mL each of UArUr, 112ArUr, and 124ArUr solutions. After achieving equilibrium, the supernatant solutions were withdrawn for measuring the residual concentrations of NH4+ and PO43– by ammonia electrode and ICPMS, respectively. In all three cases, almost 80% adsorption was observed. The adsorption of NH4+ and PO43– observed from ArUr was slightly lower than those obtained from NH4Cl and Na2HPO4, respectively. The difference is attributable to the presence of numerous other cations and anions in ArUr, which can compete with NH4+ and PO43– adsorption.

3.5. Comparison of NH4+ and PO43– Adsorption

We performed an elaborate literature study for the adsorption of NH4+ and PO43– by various reported natural polymers, agricultural waste, zeolites, clays, and chemically modified inorganic materials at varying initial concentrations (i.e., 20–1000/1–60 ppm for NH4+/PO43–), temperatures (i.e., 294–323 K), and pHo (i.e., 3–8). This survey is depicted in Table 6. Based on the comparison, the maximum adsorption capacities of DCBP/Ch are the highest with respect to the previously reported adsorbents.

Table 6. Comparison of Adsorbents.

Adsorbate Name of adsorbents Adsorption capacities (mg g–1)/pHo/C0 (ppm)/temperature (K) refs
NH4+ LMWCa 0.769/7.00/50/303 (35)
HMWCb 0.331/7.00/50/303 (35)
NZc 2.162/7.00/50/303 (35)
ANZd 2.937/7.00/50/303 (35)
Natural Turkish (Yıldızeli) zeolite 9.64/8.00/20–400/294 (36)
Natural Chinese zeolite 9.41/8.00/80/300 (37)
Transcarpathian clinoptilolite 11.6/7.00/1000/298 (38)
Natural Turkish zeolite 8.12/7.00/25–150/298 (39)
Orange peel biochar 4.71/9.00/40/298 (40)
Pineapple peel biochar 5.60/9.00/40/298 (40)
Pitaya peel biochar 2.65/9.00/40/298 (40)
Pomegranate peel 6.18/4.00/30/298 (41)
BNPPe 34.48/7.00/60/293–313 (42)
DCBP/Chf 42.16/7.00/5–30/298 Present study
PO43– TBg 1.540/7.00/1/298 (43)
CTBh 3.608/7.00/1/298 (43)
FTBi 5.434/7.00/1/298 (43)
MOP700j 1.24/7.00/2.4/298 (44)
Zenith/Fek 11.15/7.00/50/298 (45)
bentonite 4.12/7.00/50/298 (45)
Phoslock 11.60/7.00/50/298 (45)
Al-Bentl 12.70/3.00/25–60/303–323 (46)
Fe-Bentm 11.20/3.00/25–60/303–323 (46)
Fe–Al-Bentn 10.50/3.00/25–60/303–323 (46)
DCBP/Chf 15.91/4.00/1–10/298 Present study
a

Low molecular weight chitosan.

b

High molecular weight chitosan.

c

Natural zeolite.

d

Activated natural zeolite.

e

Banana peel powder.

f

dried and crushed banana peel/chitosan composite.

g

Tangerine peel biochars.

h

Tangerine peel biochars chemically activated by CaCl2.

i

Tangerine peel biochars chemically activated by FeCl3.

j

Orange peel and magnetite incorporated magnetic biochar.

k

Modified inorganic bentonite.

l

Hydroxy-aluminum pillared bentonite.

m

Hydroxy-iron pillared bentonite.

n

Mixed hydroxy-iron–aluminum pillared bentonite.

4. Conclusion

The persistent worldwide use of synthetic fertilizers, mostly containing NH4+ and PO43–, is accompanied by enormous energy burden and environmental disruption. Human urine and industrial effluents contain significant amounts of NH4+ and PO43– ions. Harvesting nitrogen and phosphorus from such sources could lower the reliance on energy-intensive processes while simultaneously addressing environmental pollution. We have previously established the utility of banana waste via dried peels toward effective NH4+ adsorption. In this work, we have developed a pH-responsive biocomposite, DCBP/Ch, by grafting of chitosan with DCBP that is suitable for capturing NH4+ and PO43–. A combination of 1H NMR, FTIR, XRD, FESEM, EDX, and TGA suggests successful grafting of chitosan onto DCBP, resulting in distinctive structural and surface characteristics and thermal properties. The adsorption of NH4+ and PO43– on DCBP/Ch is suitably modeled by pseudo-second-order kinetics that indicates chemisorption of both ionic species. DCBP/Ch displays excellent adsorption capacities for NH4+ and PO43–, respectively. The biocomposite retains a handsome adsorptive capacity when tested on artificial urine, albeit competing with a host of other ions that are present therein. To the best of our knowledge, this is the first report of a biocomposite adsorbent that is capable of harvesting both NH4+ and PO43–. Considering the ease of preparation, inexpensive components, environmental compatibility, and excellent adsorption capacity, DCBP/Ch can be viewed as an attractive prospect for accessing NH4+ and PO43– from waste channels. The construction and detailed characterization of DCBP/Ch reported in this work also serve as instructive components for further refinement and deployment and are currently being pursued in our laboratory.

Acknowledgments

The authors are grateful to Central Instrumental Facility (CIF) and DSIR-IITGN Common Research & Technology Development Hub (CRTDH) of IIT Gandhinagar for their support towards this work. H.M. is grateful to IIT Gandhinagar for a postdoctoral fellowship. The authors gratefully acknowledge financial support for this work by GUJCOST vide project no. GUJCOST/STI/2023-24/254.

Glossary

Abbreviation

AC

adsorption capacity

α-l-Araf

α-l-arabinofuranose

ArUr

artificial urine

asym.

asymmetric

BC

bidentate chelation mode

BNPP

banana peel powder

CA

catecholamines

DCBP

dried and crushed banana peel

DCBP/Ch

biocomposite of DCBP and chitosan

def.

deformation

DLS

dynamic light scattering

EDX

energy dispersive X-ray analysis

F

flavonols

FESEM

field emission scanning electron microscope

Fols

flavan-3-ols

FTIR

Fourier-transform infrared spectroscopy

α-d-GalA

α-d-galacturonic acid

β-d-Galp

β-d-galactopyranose

d-GlcA

4-O-methyl-d-glucuronic acid

HCA

hydroxycinnamic acids

I

ionic mode

ICPMS

inductively coupled plasma mass spectrometry

Leu

leucine

MQW

milli-Q-water

N-DCBP/Ch

NH4+-adsorbed biocomposite

NMR

nuclear magnetic resonance spectroscopy

P-DCBP/Ch

PO43–-adsorbed biocomposite

Phe

phenylalanine

pHo

operating pH

pHPZC

pH at point of zero charge

α-l-Rhap

α-l-Rhamnopyranose

str.

stretching

sym.

symmetric

TGA

thermogravimetric analysis

Thr

threonine

UV–vis

ultraviolet–visible spectroscopy

Val

valine

WHO

world health organization

XRD

X-ray diffractometry

β-d-Xylp

β-d-Xylopyranose

λmax

wavelength of the maximum absorbance

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c05229.

  • Composition of artificial urine; TGA analyses of DCBP, chitosan, DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch; 1H NMR of chitosan; 1H NMR of N-DCBP/Ch; 1H NMR of P-DCBP/Ch; DLS study of DCBP, DCBP/Ch, N-DCBP/Ch, and P-DCBP/Ch; UV–vis analyses of DCBP and DCBP/Ch; pHpzc analyses of DCBP and DCBP/Ch; linear isotherm and linear kinetics fitting of NH4+ and PO43– adsorption; activation energy analyses of NH4+ and PO43– adsorption; adsorption thermodynamics fitting of PO43– adsorption; and diffusion studies of NH4+ and PO43– adsorption (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao3c05229_si_001.pdf (999.8KB, pdf)

References

  1. Hilton S. P.; Keoleian G. A.; Daigger G. T.; Zhou B.; Love N. G. Life Cycle Assessment of Urine Diversion and Conversion to Fertilizer Products at the City Scale. Environ. Sci. Technol. 2021, 55, 593–603. 10.1021/acs.est.0c04195. [DOI] [PubMed] [Google Scholar]
  2. Cespi D.; Passarini F.; Vassura I.; Cavani F. Butadiene from Biomass, a Life Cycle Perspective to Address Sustainability in the Chemical Industry. Green Chem. 2016, 18, 1625–1638. 10.1039/C5GC02148K. [DOI] [Google Scholar]
  3. Chojnacka K.; Moustakas K.; Witek-Krowiak A. Bio-Based Fertilizers: A Practical Approach Towards Circular Economy. Bioresour. Technol. 2020, 295, 122223 10.1016/j.biortech.2019.122223. [DOI] [PubMed] [Google Scholar]
  4. Svanbäck A.; McCrackin M. L.; Swaney D. P.; Linefur H.; Gustafsson B. G.; Howarth R. W.; Humborg C. Reducing Agricultural Nutrient Surpluses in a Large Catchment – Links to Livestock Density. Sci. Total Environ. 2019, 648, 1549–1559. 10.1016/j.scitotenv.2018.08.194. [DOI] [PubMed] [Google Scholar]
  5. Hussain A.; Memon J. A.; Murshed M.; Alam M. S.; Mehmood U.; Alam M. N.; Rahman M.; Hayat U. A Time Series Forecasting Analysis of Overall and Sector-Based Natural Gas Demand: A Developing South Asian Economy Case. Environ. Sci. Pollut. Res. 2022, 29, 71676–71687. 10.1007/s11356-022-20861-3. [DOI] [PubMed] [Google Scholar]
  6. Erbil A. C.; Soyer E.; Beler Baykal B. Ammonium Ion Removal with a Natural Zeolite in Monodispersed and Segregated Fluidized Beds. Ind. Eng. Chem. Res. 2011, 50, 6391–6403. 10.1021/ie1014519. [DOI] [Google Scholar]
  7. Regmi U.; Boyer T. H. Ammonium and Potassium Removal from Undiluted and Diluted Hydrolyzed Urine Using Natural Zeolites. Chemosphere 2021, 268, 128849 10.1016/j.chemosphere.2020.128849. [DOI] [PubMed] [Google Scholar]
  8. Sarigul N.; Korkmaz F.; Kurultak İ. A New Artificial Urine Protocol to Better Imitate Human Urine. Sci. Rep. 2019, 9, 20159 10.1038/s41598-019-56693-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Pathy A.; Ray J.; Paramasivan B. Challenges and Opportunities of Nutrient Recovery from Human Urine Using Biochar for Fertilizer Applications. J. Cleaner Prod. 2021, 304, 127019 10.1016/j.jclepro.2021.127019. [DOI] [Google Scholar]
  10. Yapicioglu A.; Dincer I. A Review on Clean Ammonia as a Potential Fuel for Power Generators. Renewable Sustainable Energy Rev. 2019, 103, 96–108. 10.1016/j.rser.2018.12.023. [DOI] [Google Scholar]
  11. Ray H.; Perreault F.; Boyer T. H. Ammonia Recovery from Hydrolyzed Human Urine by Forward Osmosis with Acidified Draw Solution. Environ. Sci. Technol. 2020, 54, 11556–11565. 10.1021/acs.est.0c02751. [DOI] [PubMed] [Google Scholar]
  12. Elser J. J.; Bracken M. E. S.; Cleland E. E.; Gruner D. S.; Harpole W. S.; Hillebrand H.; Ngai J. T.; Seabloom E. W.; Shurin J. B.; Smith J. E. Global Analysis of Nitrogen and Phosphorus Limitation of Primary Producers in Freshwater, Marine and Terrestrial Ecosystems. Ecol. Lett. 2007, 10, 1135–1142. 10.1111/j.1461-0248.2007.01113.x. [DOI] [PubMed] [Google Scholar]
  13. Vitousek P. M.; Aber J. D.; Howarth R. W.; Likens G. E.; Matson P. A.; Schindler D. W.; Schlesinger W. H.; Tilman D. G. Human Alteration of the Global Nitrogen Cycle: Sources and Consequences. Ecol. Appl. 1997, 7, 737–750. 10.1890/1051-0761(1997)007[0737:HAOTGN]2.0.CO;2. [DOI] [Google Scholar]
  14. Ghaly A. E.; Ramakrishnan V. V. Nitrogen Sources and Cycling in the Ecosystem and its Role in Air, Water and Soil Pollution: A Critical Review. J. Pollut. Eff. Cont. 2015, 3, 136. 10.4172/2375-4397.1000136. [DOI] [Google Scholar]
  15. Schrader F.; Brümmer C. Land Use Specific Ammonia Deposition Velocities: A review of recent studies (2004–2013). Water Air Soil Pollut. 2014, 225, 2114. 10.1007/s11270-014-2114-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Zahed M. A.; Salehi S.; Tabari Y.; Farraji H.; Ataei-Kachooei S.; Zinatizadeh A. A.; Kamali N.; Mahjouri M. Phosphorus Removal and Recovery: State of the Science and Challenges. Environ. Sci. Pollut. Res. 2022, 29, 58561–58589. 10.1007/s11356-022-21637-5. [DOI] [PubMed] [Google Scholar]
  17. Peng L.; Dai H.; Wu Y.; Peng Y.; Lu X. A Comprehensive Review of the Available Media and Approaches for Phosphorus Recovery from Wastewater. Water Air Soil Pollut. 2018, 229, 115. 10.1007/s11270-018-3706-4. [DOI] [Google Scholar]
  18. Liu H.; Dong Y.; Liu Y.; Wang H. Screening of Novel Low-Cost Adsorbents from Agricultural Residues to Remove Ammonia Nitrogen from Aqueous Solution. J. Hazard. Mater. 2010, 178, 1132–1136. 10.1016/j.jhazmat.2010.01.117. [DOI] [PubMed] [Google Scholar]
  19. Huang J.; Xu C.; Ridoutt B. G.; Wang X.; Ren P. Nitrogen and Phosphorus Losses and Eutrophication Potential Associated with Fertilizer Application to Cropland in China. J. Clean. Prod. 2017, 159, 171–179. 10.1016/j.jclepro.2017.05.008. [DOI] [Google Scholar]
  20. Pradhan S. K.; Mikola A.; Vahala R. Nitrogen and Phosphorus Harvesting from Human Urine Using a Stripping, Absorption, and Precipitation Process. Environ. Sci. Technol. 2017, 51, 5165–5171. 10.1021/acs.est.6b05402. [DOI] [PubMed] [Google Scholar]
  21. Bonmatí A.; Flotats X. Air Stripping of Ammonia from Pig Slurry: Characterisation and Feasibility as a Pre- or Post-Treatment to Mesophilic Anaerobic Digestion. Waste Manage. 2003, 23, 261–272. 10.1016/S0956-053X(02)00144-7. [DOI] [PubMed] [Google Scholar]
  22. Park S.; Kim M. Innovative Ammonia Stripping with an Electrolyzed Water System as Pretreatment of Thermally Hydrolyzed Wasted Sludge for Anaerobic Digestion. Water Res. 2015, 68, 580–588. 10.1016/j.watres.2014.10.033. [DOI] [PubMed] [Google Scholar]
  23. He L.; Xu W.; Song Y.; Liu X.; Zhao Z. Selective Removal of Magnesium from a Lithium-Concentrated Anolyte by Magnesium Ammonium Phosphate Precipitation. Sep. Purif. Technol. 2017, 187, 214–220. 10.1016/j.seppur.2017.04.028. [DOI] [Google Scholar]
  24. Tünay O.; Kabdasli I.; Orhon D.; Kolçak S. Ammonia Removal by Magnesium Ammonium Phosphate Precipitation in Industrial Wastewaters. Water Sci. Technol. 1997, 36, 225–228. 10.2166/wst.1997.0524. [DOI] [Google Scholar]
  25. Karak T.; Bhattacharyya P. Human Urine as a Source of Alternative Natural Fertilizer in Agriculture: A Flight of Fancy or an Achievable Reality. Resour. Conserv. Recycl. 2011, 55, 400–408. 10.1016/j.resconrec.2010.12.008. [DOI] [Google Scholar]
  26. Yan L.; Zhang S.; Hao G.; Zhang X.; Ren Y.; Wen Y.; Guo Y.; Zhang Y. Simultaneous Nitrification and Denitrification by EPSs in Aerobic Granular Sludge Enhanced Nitrogen Removal of Ammonium-Nitrogen-Rich Wastewater. Bioresour. Technol. 2016, 202, 101–106. 10.1016/j.biortech.2015.11.088. [DOI] [PubMed] [Google Scholar]
  27. Zhou B.; Zhang N.; Wu Y.; Yang W.; Lu Y.; Wang Y.; Wang S. An Option for Green and Sustainable Future: Electrochemical Conversion of Ammonia into Nitrogen. J. Energy Chem. 2021, 60, 384–402. 10.1016/j.jechem.2021.01.011. [DOI] [Google Scholar]
  28. Kim K. W.; Kim Y. J.; Kim I. T.; Park G. I.; Lee E. H. Electrochemical Conversion Characteristics of Ammonia to Nitrogen. Water Res. 2006, 40, 1431–1441. 10.1016/j.watres.2006.01.042. [DOI] [PubMed] [Google Scholar]
  29. Yang Z.; Liu H.; Li J.; Yang K.; Zhang Z.; Chen F.; Wang B. High-Throughput Metal Trap: Sulfhydryl-Functionalized Wood Membrane Stacks for Rapid and Highly Efficient Heavy Metal Ion Removal. ACS Appl. Mater. Interfaces 2020, 12, 15002–15011. 10.1021/acsami.9b19734. [DOI] [PubMed] [Google Scholar]
  30. Rashid T. U.; Kabir S. M. F.; Biswas M. C.; Bhuiyan M. A. R. Sustainable Wastewater Treatment via Dye-Surfactant Interaction: A Critical Review. Ind. Eng. Chem. Res. 2020, 59, 9719–9745. 10.1021/acs.iecr.0c00676. [DOI] [Google Scholar]
  31. Singha N. R.; Karmakar M.; Mahapatra M.; Mondal H.; Dutta A.; Deb M.; Mitra M.; Roy C.; Chattopadhyay P. K. An In Situ Approach for the Synthesis of a Gum Ghatti-g-Interpenetrating Terpolymer Network Hydrogel for the High-Performance Adsorption Mechanism Evaluation of Cd(II), Pb(II), Bi(III) and Sb(III). J. Mater. Chem. A 2018, 6, 8078–8100. 10.1039/C8TA01106K. [DOI] [Google Scholar]
  32. Singha N. R.; Karmakar M.; Mahapatra M.; Mondal H.; Dutta A.; Roy C.; Chattopadhyay P. K. Systematic Synthesis of Pectin-g-(Sodium Acrylate-co-N-Isopropylacrylamide) Interpenetrating Polymer Network for Superadsorption of Dyes/M(II): Determination of Physicochemical Changes in Loaded Hydrogels. Polym. Chem. 2017, 8, 3211–3237. 10.1039/C7PY00316A. [DOI] [Google Scholar]
  33. Mondal H.; Karmakar M.; Chattopadhyay P. K.; Singha N. R. Starch-g-Tetrapolymer Hydrogel via In Situ Attached Monomers for Removals of Bi(III) and/or Hg(II) and Dye(s): RSM-Based Optimization. Carbohydr. Polym. 2019, 213, 428–440. 10.1016/j.carbpol.2019.02.035. [DOI] [PubMed] [Google Scholar]
  34. Mondal H.; Karmakar M.; Chattopadhyay P. K.; Singha N. R. Synthesis of pH Responsive Sodium Alginate-g-Tetrapolymers via N–C and O–C Coupled In Situ Monomers: A Reusable Optimum Hydrogel for Removal of Plant Stressors. J. Mol. Liq. 2020, 319, 114097 10.1016/j.molliq.2020.114097. [DOI] [Google Scholar]
  35. Safie N. N.; Zahrim Yaser A.; Hilal N. Ammonium Ion Removal Using Activated Zeolite and Chitosan. Asia-Pac. J. Chem. Eng. 2020, 15, e2448 10.1002/apj.2448. [DOI] [Google Scholar]
  36. Saltali K.; Sari A.; Aydin M. Removal of Ammonium Ion from Aqueous Solution by Natural Turkish (Yıldızeli) Zeolite for Environmental Quality. J. Hazard. Mater. 2007, 141, 258–263. 10.1016/j.jhazmat.2006.06.124. [DOI] [PubMed] [Google Scholar]
  37. Huang H.; Xiao X.; Yan B.; Yang L. Ammonium Removal from Aqueous Solutions by Using Natural Chinese (Chende) Zeolite as Adsorbent. J. Hazard. Mater. 2010, 175, 247–252. 10.1016/j.jhazmat.2009.09.156. [DOI] [PubMed] [Google Scholar]
  38. Lebedynets M.; Sprynskyy M.; Sakhnyuk I.; Zbytniewski R.; Golembiewski R.; Buszewski B. Adsorption of Ammonium Ions onto a Natural Zeolite: Transcarpathian Clinoptilolite. Adsorpt. Sci. Technol. 2004, 22, 731–741. 10.1260/0263617043026541. [DOI] [Google Scholar]
  39. Karadag D.; Koc Y.; Turan M.; Armagan B. Removal of Ammonium Ion from Aqueous Solution Using Natural Turkish Clinoptilolite. J. Hazard. Mater. 2006, 136, 604–609. 10.1016/j.jhazmat.2005.12.042. [DOI] [PubMed] [Google Scholar]
  40. Hu X.; Zhang X.; Ngo H. H.; Guo W.; Wen H.; Li C.; Zhang Y.; Ma C. Comparison Study on the Ammonium Adsorption of the Biochars Derived from Different Kinds of Fruit Peel. Sci. Total Environ. 2020, 707, 135544 10.1016/j.scitotenv.2019.135544. [DOI] [PubMed] [Google Scholar]
  41. Bellahsen N.; Varga G.; Halyag N.; Kertész S.; Tombácz E.; Hodúr C. Pomegranate Peel as a New Low-Cost Adsorbent for Ammonium Removal. Int. J. Environ. Sci. Technol. 2021, 18, 711–722. 10.1007/s13762-020-02863-1. [DOI] [Google Scholar]
  42. Singh R.; Datta B. Banana Peel Powder as an Effective Multilayer Adsorbent of Ammonium Ions. Ind. Eng. Chem. Res. 2022, 61, 18464–18474. 10.1021/acs.iecr.2c03052. [DOI] [Google Scholar]
  43. Son C.; An W.; Lee G.; Jeong I.; Lee Y. G.; Chon K. Adsorption Characteristics of Phosphate Ions by Pristine, CaCl2 and FeCl3-Activated Biochars Originated from Tangerine Peels. Separations 2021, 8, 32. 10.3390/separations8030032. [DOI] [Google Scholar]
  44. Chen B.; Chen Z.; Lv S. A Novel Magnetic Biochar Efficiently Sorbs Organic Pollutants and Phosphate. Bioresour. Technol. 2011, 102, 716–723. 10.1016/j.biortech.2010.08.067. [DOI] [PubMed] [Google Scholar]
  45. Zamparas M.; Gianni A.; Stathi P.; Deligiannakis Y.; Zacharias I. Removal of Phosphate from Natural Waters Using Innovative Modified Bentonites. Appl. Clay Sci. 2012, 62–63, 101–106. 10.1016/j.clay.2012.04.020. [DOI] [Google Scholar]
  46. Yan L. G.; Xu Y. Y.; Yu H. Q.; Xin X. D.; Wei Q.; Du B. Adsorption of Phosphate from Aqueous Solution by Hydroxy-Aluminum, Hydroxy-Iron and Hydroxy-Iron–Aluminum Pillared Bentonites. J. Hazard. Mater. 2010, 179, 244–250. 10.1016/j.jhazmat.2010.02.086. [DOI] [PubMed] [Google Scholar]
  47. Ahmad T.; Danish M. Prospects of Banana Waste Utilization in Wastewater Treatment: A Review. J. Environ. Manage. 2018, 206, 330–348. 10.1016/j.jenvman.2017.10.061. [DOI] [PubMed] [Google Scholar]
  48. Hikal W. M.; Said-Al Ahl H. A. H.; Bratovcic A.; Tkachenko K. G.; Sharifi-Rad J.; Kacaniova M.; Elhourri M.; Atanassova M. Banana Peels: A Waste Treasure for Human Being. Hindawi, Evidence-Based Complementary and Alternative Medicine 2022, 2022, 7616452 10.1155/2022/7616452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Li X.; Tang Y.; Xuan Z.; Liu Y.; Luo F. Study on the Preparation of Orange Peel Cellulose Adsorbents and Biosorption of Cd2+ from Aqueous Solution. Sep. Purif. Technol. 2007, 55, 69–75. 10.1016/j.seppur.2006.10.025. [DOI] [Google Scholar]
  50. Pereira M. A. F.; Cesca K.; Poletto P.; de Oliveira D. New Perspectives for Banana Peel Polysaccharides and Their Conversion to Oligosaccharides. Food Res. Int. 2021, 149, 110706 10.1016/j.foodres.2021.110706. [DOI] [PubMed] [Google Scholar]
  51. Vu H. T.; Scarlett C. J.; Vuong Q. V. Phenolic Compounds within Banana Peel and Their Potential Uses: A Review. J. Funct. Foods 2018, 40, 238–248. 10.1016/j.jff.2017.11.006. [DOI] [Google Scholar]
  52. Gomes S.; Vieira B.; Barbosa C.; Pinheiro R. Evaluation of Mature Banana Peel Flour on Physical, Chemical, and Texture Properties of a Gluten-Free Rissol. J. Food Process. Preserv. 2022, 46, e14441 10.1111/jfpp.14441. [DOI] [Google Scholar]
  53. Zhou Y.; Gao B.; Zimmerman A. R.; Fang J.; Sun Y.; Cao X. Sorption of Heavy Metals on Chitosan-Modified Biochars and its Biological Effects. Chem. Eng. J. 2013, 231, 512–518. 10.1016/j.cej.2013.07.036. [DOI] [Google Scholar]
  54. Gao N.; Du W.; Zhang M.; Ling G.; Zhang P. Chitosan-Modified Biochar: Preparation, Modifications, Mechanisms and Applications. Int. J. Biol. Macromol. 2022, 209, 31–49. 10.1016/j.ijbiomac.2022.04.006. [DOI] [PubMed] [Google Scholar]
  55. Tan Y.; Wan X.; Ni X.; Wang L.; Zhou T.; Sun H.; Wang N.; Yin X. Efficient Removal of Cd (II) from Aqueous Solution by Chitosan Modified Kiwi Branch Biochar. Chemosphere 2022, 289, 133251 10.1016/j.chemosphere.2021.133251. [DOI] [PubMed] [Google Scholar]
  56. Lavertu M.; Xia Z.; Serreqi A. N.; Berrada M.; Rodrigues A.; Wang D.; Buschmann M. D.; Gupta A. A Validated 1H NMR Method for the Determination of the Degree of Deacetylation of Chitosan. J. Pharm. Biomed. Anal. 2003, 32, 1149–1158. 10.1016/S0731-7085(03)00155-9. [DOI] [PubMed] [Google Scholar]
  57. El-banna F. S.; Mahfouz M. E.; Leporatti S.; El-Kemary M.; Hanafy N. A. N. Chitosan as a Natural Copolymer with Unique Properties for the Development of Hydrogels. Appl. Sci. 2019, 9, 2193. 10.3390/app9112193. [DOI] [Google Scholar]
  58. Aranaz I.; Alcantara A. R.; Civera M. C.; Arias C.; Elorza B.; Heras Caballero A.; Acosta N. Chitosan: An Overview of its Properties and Applications. Polymers 2021, 13, 3256. 10.3390/polym13193256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Domard A. pH and c.d. Measurements on a Fully Deacetylated Chitosan: Application to Cu(II)-Polymer Interactions. Int. J. Biol. Macromol. 1987, 9, 98–104. 10.1016/0141-8130(87)90033-X. [DOI] [Google Scholar]
  60. Rinaudc M.; Pavlov G.; Desbrieres J. Solubilization of Chitosan in Strong Acid Medium. Int. J. Polym. Anal. Charact. 1999, 5, 267–276. 10.1080/10236669908009742. [DOI] [Google Scholar]
  61. Schulze G.; Liu C. Y.; Brodowski M.; Elsholz O.; Frenzel W.; Möller J. Different Approaches to the Determination of Ammonium Ions at Low Levels by Flow Injection Analysis. Anal. Chim. Acta 1988, 214, 121–136. 10.1016/S0003-2670(00)80435-3. [DOI] [Google Scholar]
  62. Pröfrock D.; Leonhard P.; Wilbur S.; Prange A. Sensitive, Simultaneous Determination of P, S, Cl, Br and I Containing Pesticides in Environmental Samples by GC Hyphenated with Collision-Cell ICP-MS. J. Anal. At. Spectrom. 2004, 19, 623–631. 10.1039/B310530J. [DOI] [Google Scholar]
  63. Zango Z. U. Cationic Dyes Removal Using Low-Cost Banana Peel Biosorbent. Am. J. Mater. Sci. 2018, 8, 32–38. 10.5923/j.materials.20180802.02. [DOI] [Google Scholar]
  64. Karmakar M.; Mondal H.; Ghosh T.; Chattopadhyay P. K.; Maiti D. K.; Singha N. R. Chitosan-Grafted Tetrapolymer Using Two Monomers: pH-Responsive High-Performance Removals of Cu(II), Cd(II), Pb(II), Dichromate, and Biphosphate and Analyses of Adsorbed Microstructures. Environ. Res. 2019, 179, 108839 10.1016/j.envres.2019.108839. [DOI] [PubMed] [Google Scholar]
  65. Happi Emaga T.; Robert C.; Ronkart S. N.; Wathelet B.; Paquot M. Dietary Fibre Components and Pectin Chemical Features of Peels During Ripening in Banana and Plantain Varieties. Bioresour. Technol. 2008, 99, 4346–4354. 10.1016/j.biortech.2007.08.030. [DOI] [PubMed] [Google Scholar]
  66. Rehinan Z. U.; Rashid M.; Shah W. H. Insoluble Dietary Fibre Components of Food Legumes as Affected by Soaking and Cooking Processes. Food Chem. 2004, 85, 245–249. 10.1016/j.foodchem.2003.07.005. [DOI] [Google Scholar]
  67. Lawrie G.; Keen I.; Drew B.; Chandler-Temple A.; Rintoul L.; Fredericks P.; Grøndahl L. Interactions between Alginate and Chitosan Biopolymers Characterized Using FTIR and XPS. Biomacromolecules 2007, 8, 2533–2541. 10.1021/bm070014y. [DOI] [PubMed] [Google Scholar]
  68. Munagapati V. S.; Yarramuthi V.; Kim Y.; Lee K. M.; Kim D. S. Removal of Anionic Dyes (Reactive Black 5 and Congo Red) from Aqueous Solutions Using Banana Peel Powder as an Adsorbent. Ecotoxicol. Environ. Saf. 2018, 148, 601–607. 10.1016/j.ecoenv.2017.10.075. [DOI] [PubMed] [Google Scholar]
  69. Dmochowska A.; Czajkowska J.; Jędrzejewski R.; Stawiński W.; Migdał P.; Fiedot-Toboła M. Pectin Based Banana Peel Extract as a Stabilizing Agent in Zinc Oxide Nanoparticles Synthesis. Int. J. Biol. Macromol. 2020, 165, 1581–1592. 10.1016/j.ijbiomac.2020.10.042. [DOI] [PubMed] [Google Scholar]
  70. Socrates G.In Infrared and Raman Characteristic Group Frequencies, 3rd ed.; John Wiley & Sons Ltd.: Sussex, England, 2001; pp 50–134. [Google Scholar]
  71. Mondal H.; Karmakar M.; Ghosh N. N.; Maiti D. K.; Chattopadhyay P. K.; Singha N. R. One-Pot Synthesis of Sodium Alginate-Grafted-Terpolymer Hydrogel for As(III) and V(V) Removal: In Situ Anchored Comonomer and DFT Studies on Structures. J. Environ. Manage. 2021, 294, 112932 10.1016/j.jenvman.2021.112932. [DOI] [PubMed] [Google Scholar]
  72. Yang H.; Shi B.; Xue Y.; Ma Z.; Liu L.; Liu L.; Yu Y.; Zhang Z.; Annamalai P. K.; Song P. Molecularly Engineered Lignin-Derived Additives Enable Fire-Retardant, UV-Shielding, and Mechanically Strong Polylactide Biocomposites. Biomacromolecules 2021, 22, 1432–1444. 10.1021/acs.biomac.0c01656. [DOI] [PubMed] [Google Scholar]
  73. Durkin D. P.; Hoffman C. E.; Gulbrandson A. J.; Fessler E. R.; Chung R. T.; Aiello A.; Cosby T.; Woodcock J. W.; Trulove P. C. Fiber-Welded Polyionic Biocomposites Using 1-alkyl-3-vinylimidazolium Alkylphosphonate Ionic Liquids. J. Ionic Liq. 2022, 2, 100024 10.1016/j.jil.2022.100024. [DOI] [Google Scholar]
  74. Salama H. E.; Abdel Aziz M. S. Non-Toxic Chitosan-Pyrazole Adsorbent Enriched with Greenly Synthesized Zinc Oxide Nanoparticles for Dye Removal from Wastewater. Int. J. Biol. Macromol. 2023, 241, 124632 10.1016/j.ijbiomac.2023.124632. [DOI] [PubMed] [Google Scholar]
  75. Noguchi M.; Hasegawa Y.; Suzuki S.; Nakazawa M.; Ueda M.; Sakamoto T. Determination of Chemical Structure of Pea Pectin by Using Pectinolytic Enzymes. Carbohydr. Polym. 2020, 231, 115738 10.1016/j.carbpol.2019.115738. [DOI] [PubMed] [Google Scholar]
  76. Nagasundaram N.; Kokila M.; Sivaguru P.; Santhosh R.; Lalitha A. SO3H@Carbon Powder Derived from Waste Orange Peel: An Efficient, Nano-Sized Greener Catalyst for the Synthesis of Dihydropyrano[2,3-C]Pyrazole Derivatives. Adv. Powder Technol. 2020, 31, 1516–1528. 10.1016/j.apt.2020.01.012. [DOI] [Google Scholar]
  77. Suteris N. N.; Yasin A.; Misnon I. I.; Roslan R.; Zulkifli F. H.; Rahim M. H. A.; Venugopal J. R.; Jose R. Curcumin Loaded Waste Biomass Resourced Cellulosic Nanofiber Cloth as a Potential Scaffold for Regenerative Medicine: An In-Vitro Assessment. Int. J. Biol. Macromol. 2022, 198, 147–156. 10.1016/j.ijbiomac.2021.12.006. [DOI] [PubMed] [Google Scholar]
  78. Mondal H.; Karmakar M.; Chattopadhyay P. K.; Halder A.; Singha N. R. Scale-Up One-Pot Synthesis of Waste Collagen and Apple Pomace Pectin Incorporated Pentapolymer Biocomposites: Roles of Waste Collagen for Elevations of Properties and Unary/Ternary Removals of Ti(IV), As(V), and V (V). J. Hazard. Mater. 2021, 409, 124873 10.1016/j.jhazmat.2020.124873. [DOI] [PubMed] [Google Scholar]
  79. Karmakar M.; Mondal H.; Mahapatra M.; Chattopadhyay P. K.; Chatterjee S.; Singha N. R. Pectin-Grafted Terpolymer Superadsorbent via N–H Activated Strategic Protrusion of Monomer for Removals of Cd(II), Hg(II), and Pb(II). Carbohydr. Polym. 2019, 206, 778–791. 10.1016/j.carbpol.2018.11.032. [DOI] [PubMed] [Google Scholar]
  80. Isogai A. NMR Analysis of Cellulose Dissolved in Aqueous NaOH Solutions. Cellulose 1997, 4, 99–107. 10.1023/A:1018471419692. [DOI] [Google Scholar]
  81. Egüés I.; Sanchez C.; Mondragon I.; Labidi J. Separation and Purification of Hemicellulose by Ultrafiltration. Ind. Eng. Chem. Res. 2012, 51, 523–530. 10.1021/ie202304q. [DOI] [Google Scholar]
  82. Lan W.; Liu C.; Sun R. Fractionation of Bagasse into Cellulose, Hemicelluloses, and Lignin with Ionic Liquid Treatment Followed by Alkaline Extraction. J. Agric. Food Chem. 2011, 59, 8691–8701. 10.1021/jf201508g. [DOI] [PubMed] [Google Scholar]
  83. Sun S. N.; Yuan T. Q.; Li M. F.; Cao X. F.; Xu F.; Liu Q. Y. Structural Characterization of Hemicelluloses from Bamboo Culms (Neosinocalamus Affinis). Cellul. Chem. Technol. 2012, 46, 165–176. [Google Scholar]
  84. de Sousa Nascimento L.; da Mata Vieira F. I. D.; Horacio V.; Marques F. P.; Rosa M. F.; Souza S. A.; de Freitas R. M.; Uchoa D. E. A.; Mazzeto S. E.; Lomonaco D.; Avelino F. Tailored Organosolv Banana Peels Lignins: Improved Thermal, Antioxidant and Antimicrobial Performances by Controlling Process Parameters. Int. J. Biol. Macromol. 2021, 181, 241–252. 10.1016/j.ijbiomac.2021.03.156. [DOI] [PubMed] [Google Scholar]
  85. Mohd Zaini H.; Roslan J.; Saallah S.; Munsu E.; Sulaiman N. S.; Pindi W. Banana Peels as a Bioactive Ingredient and Its Potential Application in the Food Industry. J. Funct. Foods 2022, 92, 105054 10.1016/j.jff.2022.105054. [DOI] [Google Scholar]
  86. Pereira A. G. B.; Muniz E. C.; Hsieh Y. 1H NMR and 1H-13C HSQC Surface Characterization of Chitosan-Chitin Sheath-Core Nanowhiskers. Carbohydr. Polym. 2015, 123, 46–52. 10.1016/j.carbpol.2015.01.017. [DOI] [PubMed] [Google Scholar]
  87. Singha N. R.; Mahapatra M.; Karmakar M.; Dutta A.; Mondal H.; Chattopadhyay P. K. Synthesis of Guar Gum-g-(Acrylic Acid-co-Acrylamide-co-3-Acrylamido Propanoic Acid) IPN via In Situ Attachment of Acrylamido Propanoic Acid for Analyzing Superadsorption Mechanism of Pb(II)/Cd(II)/Cu(II)/MB/MV. Polym. Chem. 2017, 8, 6750–6777. 10.1039/C7PY01564J. [DOI] [Google Scholar]
  88. Xu J.; Butler I. S.; Gilson D. F. R. FT-Raman and High-Pressure Infrared Spectroscopic Studies of Dicalcium Phosphate Dihydrate (CaHPO4·2H2O) and Anhydrous Dicalcium Phosphate (CaHPO4). Spectrochim. Acta, Part A 1999, 55, 2801–2809. 10.1016/S1386-1425(99)00090-6. [DOI] [Google Scholar]
  89. Filippov L. O.; Kaba O. B.; Filippova I. V. Surface Analyses of Calcite Particles Reactivity in the Presence of Phosphoric Acid. Adv. Powder Technol. 2019, 30, 2117–2125. 10.1016/j.apt.2019.06.026. [DOI] [Google Scholar]
  90. Mondal H.; Karmakar M.; Chattopadhyay P. K.; Singha N. R. New Property-Performance Optimization of Scalable Alginate-g-Terpolymer for Ce(IV), Mo(VI), and W(VI) Exclusions. Carbohydr. Polym. 2020, 245, 116370 10.1016/j.carbpol.2020.116370. [DOI] [PubMed] [Google Scholar]
  91. Mondal H.; Karmakar M.; Dutta A.; Mahapatra M.; Deb M.; Mitra M.; Roy J. S. D.; Roy C.; Chattopadhyay P. K.; Singha N. R. Tetrapolymer Network Hydrogels via Gum Ghatti-Grafted and N–H/C–H-Activated Allocation of Monomers for Composition-Dependent Superadsorption of Metal Ions. ACS Omega 2018, 3, 10692–10708. 10.1021/acsomega.8b01218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Karmakar M.; Mahapatra M.; Dutta A.; Chattopadhyay P. K.; Singha N. R. Fabrication of Semisynthetic Collagenic Materials for Mere/Synergistic Adsorption: A Model Approach of Determining Dye Allocation by Systematic Characterization and Optimization. Int. J. Biol. Macromol. 2017, 102, 438–456. 10.1016/j.ijbiomac.2017.04.044. [DOI] [PubMed] [Google Scholar]
  93. López-Luna J.; Ramírez-Montes L. E.; Martinez-Vargas S.; Martínez A. I.; Mijangos-Ricardez O. F.; González-Chávez M. C. A.; Carrillo-González R.; Solís-Domínguez F. A.; Cuevas-Díaz M. C.; Vázquez-Hipólito V. Linear and Nonlinear Kinetic and Isotherm Adsorption Models for Arsenic Removal by Manganese Ferrite Nanoparticles. SN Appl. Sci. 2019, 1, 950. 10.1007/s42452-019-0977-3. [DOI] [Google Scholar]
  94. Karmakar M.; Mondal H.; Ghosh N. N.; Chattopadhyay P. K.; Singha N. R. Synthesis of Gum Tragacanth-Grafted Pentapolymer Hydrogels for As(III) Exclusion: Roles of Microwaves, RSM Optimization, and DFT Studies. Int. J. Biol. Macromol. 2021, 184, 909–925. 10.1016/j.ijbiomac.2021.06.070. [DOI] [PubMed] [Google Scholar]
  95. Bouatra S.; Aziat F.; Mandal R.; Guo A. C.; Wilson M. R.; Knox C.; Bjorndahl T. C.; Krishnamurthy R.; Saleem F.; Liu P.; Dame Z. T.; Poelzer J.; Huynh J.; Yallou F. S.; Psychogios N.; Dong E.; Bogumil R.; Roehring C.; Wishart D. S. The Human Urine Metabolome. PLoS One 2013, 8, e73076 10.1371/journal.pone.0073076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Tarpeh W. A.; Wald I.; Wiprächtiger M.; Nelson K. L. Effects of Operating and Design Parameters on Ion Exchange Columns for Nutrient Recovery from Urine. Environ. Sci.: Water Res. Technol. 2018, 4, 828–838. 10.1039/C7EW00478H. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao3c05229_si_001.pdf (999.8KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES