Abstract
In the modern era, sol-gel plays a key role in the progress of a new generation of dispersive solid-phase microextractors (d–µ SPMEs) for the removal of organic and inorganic pollutants in complex matrices. Thus, the current study reports the use of sol-gel-functionalized polyurethane foams (PUFs) as a novel solid platform for complete extraction of chromium (VI) species from aqueous media. The planned protocol was based upon the complete extraction of the formed binary complex ion associates between the protonated ether and/or urethane groups of PUFs and chlorochromate anion [CrO3Cl]−aq in aqueous HCl (≥1M) medium in addition to H-bonding and the electrostatic π–π interaction that resulted between the CrO3Cl− and the silanol group (Si/ZrO2, Si–O–Zr) and siloxane (Si–O–Si) groups of the sol-gel. The impact of the analytical parameters (solution pH, natural mineral acids, shaking time, temperature, and chromium (VI) concentrations) was critically studied. At the optimal conditions, the uptake capacity of the established extractor (9.9 mg·g−1) was in agreement with the Langmuir adsorption capacity (12.08 mg·g−1) of the monolayer. The sorption data fitted well with the pseudo first-order kinetic model (R2 = 0.9961) with an overall rate constant (k1) of 0.081 min−1 and an equilibrium capacity (qe) of 8.6 mg·g−1, which is in a good agreement with the experimental value (9.9 mg·g−1). The sorption of the oxyion [CrO3Cl]−aq onto the solid sorbent is an endothermic and spontaneous process as reflected from the values of ΔH (6.99 kJ·mol−1) and ΔG (−8.14 kJ·mol−1 at 293 K), respectively. The ΔS value (15.13 kJ·mol−1·K−1) reflects that the [CrO3Cl]−aq retention onto the sol-gel-treated PUFs sorbent proceeded in a more unplanned fashion. Sol-gel-treated PUFs sorbent-packed minicolumns were successfully used for the complete removal of trace levels of chromium (VI) species from water samples. Sorbed chromium (VI) species were recovered with NaOH (0.5 M) and analysed by spectrophotometry, which supports the utility of the sol-gel-treated PUFs as a low-cost solid extractor for water treatment.
1. Introduction
Chromium exists in nature in two oxidation forms mainly chromium (III) and chromium (VI) as a liquid, solid, or gas [1] and could reach the aquatic environment from different natural sources and industrial activities [2–4]. Chromium (VI) species such as chromate and/or dichromate ions in water have genotoxic and mutagenic effects on biological systems, while trivalent chromium plays a significant role in the metabolism of humans and retains much lower toxicity [5, 6]. Thus, chromium (VI) in the aquatic environment represents one of the key contaminants; the US-EPA has classified chromium as a group “A” carcinogenic agent of anthropological toxins [4, 7]. Exposure to high levels of chromium (VI) causes breathing problems [7, 8], allergic reactions, ulceration of the skin, and damage to the eyes, liver, kidneys, circulatory and nerve tissues, as well as immune systems [9–11]. On the other hand, chromium (VI) toxicity from food remains a great concern [12]. According to US-EPA and WHO, the maximum allowable edges of chromium (VI) in drinking water are 0.1 µg/mL and 0.05 µg/mL, respectively [13].
A series of analytical techniques for trace determination and preconcentration of chromium species in particular chromate and/or dichromate has been reported [14–22]. The most common spectroscopic techniques are spectrofluorimetry, spectrophotometry [18–20, 22] atomic absorption spectrometry (AAS) [23–26], inductive coupled plasma-mass spectrometry (ICP-MS) [27, 28], atomic emission spectrometry (AES) [29, 30], and total reflection X-ray fluorescence spectrometry (TRXRFS) [31]. Moreover, chromatographic separation and determination involving high-performance liquid chromatography (HPLC) [32–37] and gas chromatography (GC) [8] have also been applied for the trace determination of chromium species in complex matrices. However, complications exist at trace and ultratrace levels with direct quantification of chromium; thus, the enrichment step is compulsory compared to traditional tools with time-consuming approaches, use of large volumes of solvents, and unsatisfactory recovery percentage [33, 34].
Nowadays, the extraordinary development in the use of dispersive solid-phase microextractors (d–µ SPMEs) has attracted many researchers [38, 39]. This tool offers a widespread range of applications besides its cost effectiveness, ease of the process, low solvent consumption, high recovery, and rapidity [38]. Solid phase extraction (SPE) and/or solid phase microextraction (SPME) represent one of the most effective techniques that are frequently applied for trace and ultratrace determination of dyes and metal ions in environmental water samples [40–43] and removing pollution from water [44], whether in batch or column (flow) mode. Sponge solid-phase extraction involving the use of polyurethane foams (PUFs) of polyether and/or polyester types as a flexible membrane-like structure has received considerable attention in the last few decades [45, 46]. PUFs have been used as an outstanding substrate for solid-phase extraction for separation and/or preconcentration procedures for trace quantification of various species such as organic and inorganic oxyanions and metal ions [45, 46]. In PUFs, the occurrence of polar and nonpolar groups enhances the analyte sorption [47, 48]. PUFs have several advantages such as chemical stability, availability, and affordability; moreover, they can easily be reused for dyes and other organic pollutants removal desorption [49, 50]. PUFs are also beneficial in column procedures in off-line or flow injection preconcentration systems, since this class of solid-phase extractors has minimum resistance for fluid passage and does not cause over pressure or swelling like other sorbents [46]. The large surface area of nanoparticles could enhance the sensitivity of the solid sorbents and modify their properties [51, 52] by coupling them together in different forms [53–57]. The sol-gel technology with the advantages of optical transparency and chemical robustness has received considerable interest in various applications in the last few years [58–60]. Hence, the current study aimed to (i) synthesize and characterize the sol-gel physically impregnated PUFs as a solid microextractor; (ii) study the retention profile of chromium (VI) from the aqueous by the established sol-gel-treated PUFs solid platform; (iii) properly assign the kinetics, thermodynamics, sorption models, and multimode sorption mechanisms of chromium (VI) by the developed sol-gel-treated PUFs; and (iv) finally, test the analytical utility of the developed microextractor in flow (packed column) and pulse modes for complete extraction and recovery of trace levels of chromium (VI) in environmental water samples.
2. Experimental Setup
2.1. Reagents, Chemicals, and Materials
All the chemicals and solvents were analytical-reagent grade (AR) and used as received. The chemicals, NaOH, HCl, H2SO4, HNO3, K2Cr2O7, K2CrO4, and organic solvents such as acetone and isopropanol were purchased from Merck company and were used as received. The precursors methyltrimethoxysilane (C4H12O3Si), methyltriethoxysilane (C7H18O3Si), tetramethyl orthosilicate (C4H12O4Si), tetraethyl orthosilicate (C8H20O4Si), and 3-(trimethoxysilyl)propyl methacrylate (C10H20O5Si) were purchased from Aldrich (Assay > 98). 3-methacryloxypropyltrimethoxysilane (MAPTMS, assay 99% in methanol, Sigma-Aldrich), zirconium (IV) n-propoxide (ZPO, assay 70% in propanol, Sigma-Aldrich, Ireland), and methacrylic acid (MAAH, C4H6O4, assay > 98%, Sigma-Aldrich, Ireland) have been used in sol-gel preparation. Low density polyethylene (LDPE) bottles, Nalgene, were used for storage of water samples. Commercial open-cell polyether-type-based PUFs, which were supplied locally in Saudi Arabia, were cut into cubes of approximately one cm3, washed with 8% HCl and deionized water until the washing solutions were free from chloride ions [61–63], and then washed with acetone to remove organic contaminates and finally dried in an oven at 80°C for 2 h. A series of standard solutions (0.005–0.01% m/v) of the ion-pairing reagent procaine hydrochloride (PQ+.Cl−) was prepared individually in deionized water. Britton–Robinson (B–R) buffer solutions of pH 2–11.7 were prepared as reported earlier [64].
2.2. Instrumentation
A scanning electron microscope (SEM; JSM-5910, JEOL) was used for recording the surface morphology of the synthesized sol-gel and sol-gel-treated polyurethane foams (PUFs). An energy-dispersive X-ray spectroscopy (EDX) was used to confirm the elemental composition. Scanning atomic force microscopy (AFM; Pacific Nanotechnology: Nano-RTM) was also used to fully characterize the surface topography and roughness of the prepared sol-gel samples. FTIR spectra of the sol-gel and sol-gel-treated PUFs were recorded using a JASCO-430 model spectrometer. PerkinElmer's inductively coupled plasma-optical emission spectroscopy (ICP-OES), Optima (California, CT, USA), was used for chromium determination at the optimum operational parameters (Electronic Supplementary Information (ESI.1, Table 1)) as a method validation. A UV-Vis spectrometer (Shimadzu UV-Vis 1800, Japan) was used for recording the electronic spectra and the absorbance (at 355 nm) of chromium (VI) solutions before and after extractions in 10.0 mm quartz cells. A corporation precision scientific mechanical shaker (Chicago, CH, USA) and a thermostatic controlled shaker (GFL-1083 model, Germany) were used in batch experiments. Chromatographic minicolumn packed extraction was performed on the solid-phase extraction manifold system of Agilent Technologies, 1200 USA. A pH meter (inoLab pH/ion level 2) and a Milli-Q Plus system (Millipore, Bedford, MA, USA) were used for pH measurements and for providing ultrapure water, respectively.
Table 1.
Adsorption capacities of different adsorbents for removal of Cr (VI) ions from water samples.
| Adsorbent/sorbent | Shaking time (min) | q e (mg·g–1) | Ref |
|---|---|---|---|
| Red mud | 1440 | 75 | Gupta et al. [65] |
| MWCNT | — | 9.5 | Tuzen and Soylak [66] |
| Pomegranate peel | — | 9.45 | Mohammed et al. [67] |
| Chitosan | — | 7.94 | Aydın and Aksoy [68] |
| Rich husk ash | 180 | 25.64 | Bhattacharya et al. [69] |
| Grafted macadamia nutshell powder | 180 | 39.21 | Ntuli and Pakade [70] |
| Banana peel | 60 | 10.42 | Parlayici and Pehlivan [71] |
| Pleurotus mutilus biomass | 180 | 15.50 | Alouache et al. [72] |
| Sol-gel/PQ+.Cl−/PUFs | 50 | 9.9 | Present study |
2.3. Preparation of Sol-gel
In this preparation, the mixture of the two hybrid precursors was used to form stable and homogenous sol-gel. The first precursor is an organically modified silicon precursor (MAPTMS), and the other one is an organically modified zirconium complex which was prepared from the complex formation of ZPO by MAAH, as reported earlier [53]. During the initial synthesis, the relative proportions of the MAPTMS : ZPO were 80 : 20 and the theoretical hydrolysis degree was 50% against the total content of reactive alkoxide groups. The synthesis process was carried out in three sequential steps. First, two simultaneous reactions were performed for 45 min. The first reaction was the prehydrolysis of MAPTMS by using an aqueous HNO3 solution (0.1 M) with a 4 : 1 ratio. MAPTMS was immiscible with water for 5 min, and then all species became miscible due to methanol production. The parallel reaction involved chelating ZPO using MAAH to block alkoxide groups and to minimize the precipitation. The second main step was slowly adding the partially hydrolysed MAPTMS to the zirconate complex. Finally, after 5 min, neutral hydrolysis of the mixture was completed by adding deionized water gently [53]. The overall preparation of the sol-gel material has been reported by MacHugh [73], and the experimental steps are depicted in ESI.2, scheme 1.
2.4. Preparation of the Sol-gel/PQ+.Cl− Functionalized PUFs Sorbent
The sol-gel-functionalized polyurethane foam (PUFs) sorbent was fabricated as follows:
The dried PUFs cubes (about 2.0 g) were shaken with the sol-gel dissolved in isopropanol (IPA, 25 mL, 5% v/v) with efficient stirring for 20 min to physically impregnate sol-gel onto the PUFs cubes as reported [62].
The sol-gel-treated PUFs (sol-gel/PUFs) cubes were separated out by decantation and pressed between two sheets of filter paper to remove any unbound sol-gel from the PUFs.
The dried sol-gel-treated PUFs (about 2.0 g) were then shaken with procaine hydrochloride (0.005–0.01% m/v) with efficient stirring for 20 min. The sol-gel-/PQ+.Cl− -treated PUFs cubes (sol-gel/PQ+.Cl−/PUFs) were separated out by decantation. Finally, the sol-gel/PQ+.Cl−/PUFs cubes were pressed between two sheets of filter paper to remove any traces of excess PQ+.Cl−, as reported earlier [62].
A schematic diagram describing the preparation of the proposed sol-gel/PQ+.Cl−/PUFs solid extractor is illustrated in Scheme 1.
Scheme 1.

A scheme describing the preparation of the proposed sol-gel/PQ+.Cl−/PUFs solid extractor.
2.5. Static Experiments
A precise mass of sol-gel/PQ+.Cl−/PUFs cubes (0.05–0.1 g) was shaken for 50 min with an aqueous solution (100 mL) of chromium (VI) at various known concentrations (0.05–50.0 µg·mL−1) in HCl (1.0 M) in a conical flask (200 mL) at 25 ± 1°C on a mechanical shaker. After equilibration, the aqueous phase was separated out by decantation, and the amount of chromium (VI) remaining in the aqueous phase was determined spectrophotometrically from its absorbance at 355 nm (λmax) versus reagent blank. The retained chromium (VI) species on the established extractor were then calculated from the difference (Ab − Af) of the absorbance before (Ab) and after (Af) extraction.
Similarly, the impact of various parameters (pH solution, shaking time, and temperature) was critically studied, and the distribution ratio (D) and extraction percentage (%E) of chromium (VI) uptake by the established extractor were determined as reported [62].
2.6. Analytical Applications
2.6.1. Extraction and Recovery of Chromium (VI) from Deionized and Environmental Tap Water were Performed by Minicolumn (Flow) Mode of Separation
A series of deionized water and prefiltered tap water samples (100 mL) adjusted to pH ≤ 1 using HCl (1.0 M, 10 mL) and spiked with known concentrations of chromium (VI) (0.01–1.0 µg/mL) were passed individually through sol-gel/PQ+.Cl−/PUFs (0.4 ± 0.002 g) packed minicolumn at 10 mL·min−1 flow rate. Complete retention of chromium (VI) took place as indicated from the absorbance at λmax 355 nm and/or ICP-OES measurements of chromium (VI) in the effluent against the reagent blank. A 2.0 mL solution of NaOH (0.5 M) was used as an eluting agent for the complete recovery of chromium (VI) from the sol-gel/PQ+.Cl−/PUFs packed column at 0.5 mL·min−1 flow rate. The absorbance at λmax 355 nm and/or ICP-OES signal intensity of the fractions of the eluate were measured against the reagent blank. The %E and D values were computed, whereas the recovery percentage (%R) was calculated by using the following equation:
| (1) |
2.6.2. Extraction and Recovery of Chromium (VI) from Environmental Water by Pulse Mode of Separation
A series of 100 mL of prefiltered tap water samples spiked with known concentrations of chromium (VI) (0.01–1.0 µg/mL) adjusted to pH ≤ 1.0 with HCl (1.0 M, 10 mL) was transferred into a conical flask (200 mL). Equal weights (0.2 ± 0.001 g) of the established sorbent sol-gel/PQ+.Cl−/PUFs were placed in the medical syringes (20 mL, capacity) as pulse columns individually. The columns were pulsated 25–30 times through chromium (VI) solutions of the tap water. The sorbed chromium (VI) was finally recovered with NaOH (20 mL, 0.5 M) after 25–30 pulses as indicated from the absorbance at 355 nm (or ICP-OES) measurements of the recovered chromium (VI). The %R of the chromium (VI) species was finally computed using equation (1).
3. Results and Discussion
The development of a dispersive solid-phase extractor involving the use of sol-gel/PQ+.Cl−/PUFs as an ideal dispersive solid-phase microextractor for chromium (VI) removal from aqueous is a great concern. Thus, chromium (VI) sorption from aqueous media of pH < 1 by sol-gel/PQ+.Cl−/PUFs as a dispersive solid-phase extractor (d–µ SPME) was the primary study. Significant chromium (VI) uptake (>75%) onto the modified PUFs sorbent was easily achieved compared to untreated PUFs (20–25%). Thus, in the next study, sol-gel-treated PUFs were successfully used as a d–µ SPME for chromium (VI) removal and subsequent recovery from the test aqueous media.
3.1. Characterization of Sol-gel
3.1.1. Surface Morphology Characterization
Scanning electron microscope (SEM) image of the prepared hybrid sol-gel (ESI.3) displayed uniform shapes of the surface morphology of the nanohybrid sol-gel as reported earlier [55, 74]. Atomic force microscopy (AFM) of the hybrid sol-gel was recorded further for initial evaluation of the surface roughness. The AFM images (ESI.4) showed surface roughness, the nanostructure of the sol-gel, as well as support the suitability of the sol-gel as a nanoparticle to be chemically impregnated on the solid support-based PUFs surface and membrane substrate [53]. Furthermore, the energy-dispersive X-ray spectroscopy (EDX) analysis of the hybrid sol-gel was performed to identify the presence of the main elements (ESI.5). The data show the formation of nanosized hybrid sol-gel and the appearance of the main elements Si, Zr, and O in uniform distribution with high homogeneity of the material.
3.1.2. Structural Characterization
The FTIR spectrum for the prepared sol-gel in the range 500–4000 cm−1 is shown in ESI.6A. The data revealed that the characteristic vibrations at 1000, 1100, and 1250 cm−1 corresponding to Si–O–Zr, Si–O–Si, and –C=O were in good agreement with the data reported earlier [53–55]; whereas the observed vibrations at 2900 and 3450 cm−1 were safely assigned to C–H and Si–OH, respectively. In the range 850–1250 cm−1 (ESI.6B), the spectrum of the sol-gel also displayed the characteristic vibrations at 900, 950, 1100–1125, and 1200 cm−1, which were safely assigned to Si–OH, Si–O–Zr, Si–O–Si, and Si–O–C, respectively [53–55]. The FTIR spectrum of the sol-gel also displayed the presence of the same chemical vibrations of the precursor and the prepared materials [53, 54]. The observed broad stretching vibration in the range of 800–1200 cm−1 was safely assigned to the silicate network which is combined of the silanol stretches (Si–OH: 890 cm−1) and Si–O–Si and Si–O–Zr vibrations (840 and 1010–1050 cm−1) and (940 cm−1), respectively [53–55], whereas the Si–O–C in the methoxy-silane group in MAPTMS could be responsible for the stretching vibration at 1170 cm−1 [53]. The vibrations at 1635 and 1533 cm−1 were assigned to the symmetric (νs) and asymmetric (νas) vibrations of the carboxylic group (COO−), respectively, [55, 56] with Δv (COO−) of 102 cm−1 in a bidentate fashion in Zr–MAAH complex.
3.2. Optimization of the Analytical Parameters
The impact of various parameters that control the analytical utility of the established sol-gel/PQ+.Cl−/PUFs solid microextractor towards chromium (VI) removal from aqueous media and sequential determination was critically studied in more detail using batch and flow modes of separation. The pH controls the surface charge of the solid-phase extractor and the analyte dissociation and/or ionization on the active sites of the adsorbent [22, 75]. Thus, the influence of pH on chromium (VI) sorption by the sol-gel-modified PUFs was critically studied in different pH solutions (pH ≈ 1.0–12.0). The results are illustrated in Figure 1 and show that the maximum chromium (VI) retention was achieved at a pH < 1.0, where chromium (VI) is highly extracted since the quantity of HCrO4− ions is directly proportional to the quantity of the acid and the complex ion associate formed on/in the sol-gel-treated PUFs [22, 75]. When growing the solution in pH > 1, chromium (VI) retention markedly decreased and reached a minimum value since the analyte is present in a highly polar form and the solid-phase is most likely negatively charged [22, 75]. At high pH ≥ 1.5, the formation of nonextractable chromium (VI) species and/or hydrolysis of the complex ion associate are most likely predominant [22, 75].
Figure 1.

Plot of chromium (VI) retention onto the established sol-gel/PQ+.Cl−/PUFs sorbent versus pH at 25 ± 1°C after 60 min shaking time.
Chromium (VI) oxyions occur as negatively charged species and are rapidly hydrolysed forming various species (neutral, anionic, or oxyanions) conditional on the pH of the aqueous [76]. At pH < 1.0, the chemical equilibria of oxychromium (VI) species can be stated as follows [75, 76]:
| (2) |
The impact of the presence of a series of mineral acids individually such as HCl, H2SO4, HNO3, and HClO4 (1.0 M) was critically studied. The sorption of oxychromium (VI) by the proposed dispersive extractor increased in the following order:
| (3) |
In HCl media, chromium (VI) is present as a chlorochromate (CrO3Cl−) anion [75, 77], which is directly proportional to the formation and extraction of the complex ion associated in the organic phase as
| (4) |
| (5) |
At a low pH and in the presence of HCl (1.0 M), the available chelating sites ether oxygen (–CH2–O–CH2–) and/or urethane nitrogen (–NH–COO–) linkages in the PUFs sorbents membrane are protonated resulting in high chromium (VI) sorption retention as follows [62]:
With the ether oxygen group of PUFs as
| (6) |
| (7) |
With the urethane nitrogen group of PUFs as
| (8) |
| (9) |
Thus, in the subsequent study, the HCl concentration in the extraction media of chromium (VI) was adopted at 1.0 M.
The impact of shaking time (0.0–90 min) on chromium (VI) uptake from aqueous media at the optimized pH by the established solid sorbent was studied. The results (ESI.7) revealed fast retention of the chromium (VI) by the established sorbent, and the equilibrium was attained within 50 min of shaking. The available active binding sites and the surface area of the microextractor sorbent are high at the early shaking time. Thus, minimizing the equilibrium time of the SPE towards chromium (VI) sorption is expected. Therefore, it can be concluded that the chromium (VI) partitioning ratio between the sol-gel-modified PUFs and the test aqueous solution at pH < 1 is high, and that the sorbent has excellent performance. Thus, a 50 min shaking time was adopted in the subsequent experiments and the results suggested the use of the sol-gel-treated PUFs as the solid-phase extractor in packed columns.
The outcome of temperature (20–50°C) of the test aqueous solution on chromium (VI) sorption by the established sol-gel/PQ+.Cl−/PUFs microextractor at the optimized pH and shaking time was studied. The distribution factor D (ESI.8) of chromium (VI) increased on growing temperature, revealing the endothermic characteristics of analyte uptake by the established dispersive solid-phase microextractor. Moreover, growing temperature may also reduce the number of water molecules available to solvate the chromium (VI) oxyions in the extraction media at the optimized pH, which would therefore be forced out of the solvent phase into the sorbent. Some free water molecules are also favourably unconfined from the hydration sheets around the chromium (VI) species on increasing temperature, resulting in enhanced extraction [22, 78].
The influence of chromium (VI) over a wide range of equilibrium concentrations (0.0–70 µg/mL) onto the established sol-gel/PQ+.Cl−/PUFs solid sorbent at the optimal parameters of pH < 1 and shaking time (50 min) was studied. The plot of D values versus their initial concentrations in the bulk aqueous solution (ESI.9) displayed a maximum D value of chromium (VI) sorption onto the established solid dispersive extractor from diluted and/or moderate chromium (VI) levels. The D value decreased on growing chromium (VI) concentration, where the solid sorbent membranes and surface area became more saturated with the sorbed oxychromium (VI) ions. The quantity of sorbed chromium (VI) on the sol-gel/PQ+.Cl−/PUFs sorbent varied more/or less linearly at a low or moderate analyte concentration in the aqueous HCl solution.
3.3. Sorption Isotherm Study of Chromium (VI)
The uptake capacity (qe) was calculated by using the following relationships:
| (10) |
where Cο and Ce are the initial and residual chromium concentrations (mg·L−1), V is the volume of a chromium (VI) solution (L) and m is the weight of the adsorbent (g). The maximum qe was found to be 9.9 mg·g−1 at 60 μg·mL−1 initial concentration. More importantly, Langmuir and Freundlich isotherm models were applied to describe the adsorption process of Cr (VI) onto the treated PUFs. The Langmuir isotherm model indicates that a monolayer of the analyte was formed on the surface of the homogenous adsorbent, which can be demonstrated as follows:
| (11) |
where Ce (mg·L−1) is the residual concentration at equilibrium and Kl (L·mg−1) and qm are the Langmuir constants which represent the adsorption-free energy and the maximum amount of capacity adsorption of the monolayer (mg·g−1) of treated PUFs. The plot of Ce/qeversusCe was linear (Figure 2) with a correlation coefficient (R2) of 0.935. The computed values of qm and Kι were 12.08 mg·g−1 and 0.568 L·mg−1, respectively. The maximum adsorption capacity (qm) of the established sol-gel/PQ+.Cl−/PUFs extractor towards chromium (VI) was favourably compared with some of the reported solid adsorbents [79, 80]. The results are summarized in Table 1. The capacity of the developed extractor is somewhat lower compared to other extractors; however, the time consumed to reach equilibrium and the cost of the developed extractor are much lower than other extractors. Thus, in the subsequent study, great attention has been oriented towards the sol-gel-modified solid-phase extractor.
Figure 2.

Plot of Ce/qeversusCe (Langmuir isotherm model) at 25 ± 1°C.
The data were further subjected to the Freundlich isotherm model, which predicts the multilayer adsorption on a heterogeneous system as
| (12) |
where Ce represents the Cr (VI) concentration at equilibrium (mg·L−1) and qe is the amount of Cr (VI) that is adsorbed onto treated PUFs (mg·g−1). Kf (mg·g−1) and n are Freundlich constants that were obtained from the linear plot of log qe versus log Ce and help in determining the capacity and intensity of the adsorption, respectively. Kf, which can be defined as the adsorption or distribution coefficient and exemplifies the quantity of chromium adsorbed onto modified PUFs for a unit equilibrium concentration, was found to be 0.897, while the calculated n value (1.65) gives the nonlinearity degree between the solution concentration and the adsorbent. If n = 1, the adsorption is linear; if n < 1, this suggests that the adsorption process is chemical; if n > 1, this means that the physical adsorption is satisfactory [81].
According to the modelling, the static adsorption capacity, calculated by applying the Langmuir model (12.08 mg·g−1), was consistent with the uptake capacity estimated using isotherm experiments (9.9 mg·g−1).
3.4. Kinetics of Chromium (VI) Retention
The main steps of SPE and/or d–μ SPME are the transfer of the analyte from the aqueous solution to the sorbent/adsorbent surface, and then the passage of the analyte species into the interior of the solid pores and membranes [82, 83]. Hence, assigning the kinetics of chlorochromate anion (CrO3Cl−) sorption by the developed sol-gel/PQ+.Cl−/PUFs microextractor is of great significance for the minimization and/or the complete removal of chromium (VI). The primary study of chromium (VI) sorption from the test aqueous HCl solution by the developed sol-gel/PQ+.Cl−/PUFs microextractor at the optimized parameters was found to be fast and dependent on shaking time as it reached equilibrium within a ∼50–60 min shaking period. The value of the half-life time (t1/2) of sorbed chromium (VI) was low (2.8 ± 0.028 min). Therefore, the data were further subjected to many kinetic models including nonlinear [50, 83] and linear models such as Weber–Morris [85, 86], pseudofirst order (Lagergren) [86, 87], pseudosecond order [88], and Elovich [89].
The impact of time was further studied by the Weber–Morris model [85] to assign whether intraparticle diffusion or film diffusion is the rate governing step in the chromium (VI) uptake by sol-gel/PQ+.Cl−/PUFs. The most common Weber–Morris model can be stated by the following equation:
| (13) |
where Rd is the rate constant of intraparticle transport (mg·g−1·min−1/2), qt is the equivalent quantity of retained analyte (mg·g−1) at a specific time t, and C (mg·g−1) is the thickness of the boundary layer, which is the intercept from the initial linear portion of the linear plot, where a higher value of C has a greater effect on the boundary layer [85]. The plot of qtversus the square root of time was found to be linear (R2 = 0.9929) up to 45 ± 1 min and deviates on intensifying shaking time (Figure 3(a)). The linear plot at the early stage does not pass through the origin, which confirms that the rate-controlling step for sol-gel/PQ+.Cl−/PUFs is not only intraparticle diffusion [87]. The fact that at the initial stage of shaking time, the diffusion rate of the chromium (VI) species on the established solid-phase extractor was high and diminished linearly on increasing shaking time with a correlation coefficient (R2) of 0.9842. Thus, it can be concluded that the rate-controlling step of chromium (VI) sorption is film diffusion at the initial stage of extraction [86, 87]. In the second stage of chromium (VI) extraction, the intraparticle diffusion decreased, indicating that the pore volumes of the sorbent are most likely exhausted [86, 87].
Figure 3.

The kinetic model plots of chromium (VI) retention at 25 ± 1°C: Weber–Morris (a) and Lagergren model (b).
The presence of different pore sizes could be a reason behind the change in Rd values of the first and subsequent stages from 0.6378 to 0.1251 mg·g−1·min−1/2, as calculated from the slopes (Figure 3(a)) [85, 86]. Depending on the high value of the constant C (1.5755 mg·g−1), it was considered that film diffusion plays a major role in the rate-controlling step in the overall chromium (VI) uptake and the intraparticle diffusion step cannot be the rate-controlling step [86, 87]. Moreover, the first stage occurred rapidly after transferring the chromium (VI) species from the extraction solution, i.e., the uptake stages include transport of the chromium (VI) species from the bulk liquid phase to the external surface of the d–µ SPME through a hydrodynamic boundary layer or film solution (film diffusion), and then the diffusion of the chromium (VI) species from the exterior of the SPE (external diffusion) as reported for anionic dyes onto hollow polymer microcapsules [87, 89].
The data were further analysed by the pseudofirst-order (Lagergren) model [86] as
| (14) |
where k1 (min−1) is the first-order rate constant for chromium (VI) sorption and qe and qt (mg·g−1) are the quantities of sorbed chromium (VI) per unit mass of the established solid extractor at equilibrium (qe) and at time t (qt), respectively. The plot of log (qe − qt) versus time was linear (Figure 3(b)) with a correlation coefficient of (R2 = 0.9961). The calculated value of the first-order rate constant k1 was 0.081 min−1. The equilibrium capacity (qe) for chromium (VI) sorption from the intercept of the linear plot of log (qe − qt) versus time was found to be equal to 8.6 mg·g−1, which is in good agreement with the experimental value (9.9 mg·g−1).
The results were additionally analysed by the pseudosecond-order model [87] as
| (15) |
where h = k2qe2 can be assumed as the initial rate constant of the sorption step and k2 is the pseudosecond-order rate constant g(mg−1·min−1). The plot of t/qtversus time for chromium (VI) uptake by the established solid extractor was found to be linear, as shown in ESI.10, with an acceptable correlation coefficient (R2 = 0.9984). The equilibrium capacity (qe) of chromium (VI) retention, as calculated from the slope of the linear plot, was found to be 6.75 mg·g−1; whereas the pseudosecond-order rate constant (k2) was found to be equal to 0.017 g (mg−1·min−1). Furthermore, the disagreement between the calculated qe value from the pseudosecond-order model (6.75 mg·g−1) and the experimental qe value (9.9 mg·g−1) indicates that the chromium (VI) sorption by the established solid microextractor does not follow the pseudosecond-order model. The data evidently display that the pseudofirst-order model is properly fitted for chromium (VI) sorption kinetics [87].
The data were further subjected to McLintock–Elovich kinetic mode [88]. This model is frequently used for systems in which the surface of the solid sorbent is heterogeneous and applicable mainly for chemisorption kinetics. Elovich kinetic mode can be stated by the following linear equation:
| (16) |
where α (mg·g−1·min−1) is the initial sorption rate and β (mg·g−1) is the desorption constant linked to the extent of the surface coverage and activation energy for chemisorption. Figure 4 shows the plot of qtversus ln t. The plot was linear at the early stage of shaking time (ln t < 4.1 min) with a correlation coefficient (R2) of 0.9918. The α and β parameters of Elovich were safely calculated from the intercept (1/β ln (α β)) and slope (1/β) of Figure 4. The computed values of the Elovich parameters β and α were found to be equal to 0.8619 mg·g−1 and 2.83 mg·g−1·min−1, respectively.
Figure 4.

Plot of the Elovich kinetic model for chromium (VI) retention onto sol-gel/PQ+.Cl−/PUFs at pH < 1 at 25 ± 1°C.
3.5. Thermodynamic Parameters of Chromium (VI) Sorption
In SPE or d–µ SPME, the impact of the changes of the thermodynamic parameters enthalpy (ΔH), entropy (ΔS), and Gibbs free energy (ΔG) is of great importance. Thus, the sorption of the chromium (VI) species from the aqueous solution by the established sol-gel/PQ+.Cl−/PUFs microextractor at a pH < 1 was studied over a wide range of temperatures (293–333 K). Considering no complex formation and/or precipitation of CrO3Cl− species took place, and the tested species is present as a neutral species at the optimized pH < 1, the thermodynamic parameters ΔH, ΔS, and ΔG of chromium (VI) sorption onto the used sorbent can be computed from the linear plot of ln KCversus 1000/T (Figure 5), where KC is the value for analyte retention depending on the fractional attainment (Fe) of the sorption process (KC = Fe/1 − Fe). The plot was linear over the entire temperature range with an acceptable correlation coefficient (R2 = 0.9985). The equilibrium constant increased on growing temperature, revealing that the chromium (VI) retention onto the established solid extractor is an endothermic process [90]. The numerical values of ΔH, ΔS, and ΔG as evaluated from the slope and intercept of the linear plot were found to be equal to 6.99 kJ·mol−1, 15.13 kJ·mol−1·K−1, and −8.14 kJ·mol−1 (at 293 K), respectively.
Figure 5.

Plot of ln Kcversus 1000/T for chromium (VI) sorption from an aqueous HCl solution (pH < 1) onto sol-gel/PQ+.Cl−/PUFs after 60 min shaking time.
Considering Van't Hoff equation, the distribution coefficient (D) of chromium (VI) retention is linked to absolute temperature (T) according to the following equation:
| (17) |
The plot of log D versus 1000/T for chromium (VI) sorption onto the solid sorbent was linear (ESI.11). The D values of chromium (VI) sorption from the aqueous media with pH < 1 onto the established solid platform increased on growing temperature, revealing the endothermic nature of the retention step [91]. The value of ΔH of chromium (VI) uptake onto the solid sorbent as calculated from ESI.11 was found to be equal to 6.97 ± 0.3 kJ·mol−1, which agrees with the data calculated from Figure 5 (6.99 kJ·mol−1), confirming the endothermic nature of chromium (VI) uptake.
The reaction follows a physical sorption mechanism since the ΔH value (6.97 ± 0.3 kJ·mol−1) is lower than the expected value of chemical sorption 10 kJ·mol−1. The sorption capacity of the established sorbent increased on growing temperature. This trend is most likely attributed to the bonds' strength between chromium (VI) and the sorbent's active sites with increasing temperature. The Growing temperature may control the physical characteristics of the sol-gel/PQ+.Cl−/PUFs sorbent and enhances the intermolecular interactions between the sorbent and analyte. At the studied temperatures, the ΔG value indicated the feasibility and spontaneous nature of chromium (VI) uptake on increasing temperature. Furthermore, the decrease in Gibbs free energy (ΔG) with increasing temperature implies that the chromium (VI) sorption process is spontaneous, endothermic and is more favourable at a high temperature. The energy of the active sites of the sorbent provided by raising the temperature most likely enhances the possible interaction between the active sites of the established sol-gel/PQ+.Cl−/PUFs and the oxychromium (VI) ions resulting in a higher sorption percentage of the analyte.
3.6. Mechanism of Chromium (VI) Sorption
Based on the sorption characteristics, kinetics, and thermodynamic parameters of chromium (VI) uptake by the established sol-gel/PQ+.Cl−/PUFs sorbent from an aqueous solution of pH < 1, it can be recognized that the CrO3Cl− uptake is most likely attributed to a multimode sorption mechanism involving (i) absorption related to “weak anion exchange” and/or “solvent extraction” of the binary complex ion associates {[–CH2–OH+–CH2–].[CrO3Cl]−}sol-gel/PUFs, {[–NH2 + –COO–].[CrO3Cl]−}sol-gel/PUFs, and{[PQ]+.[CrO3Cl]−}sol-gel/PUFs; (ii) H-bonding between the silanol group of the sol-gel/PUFs and the bulky anion [CrO3Cl]−; and (iii) the electrostatic π–π interaction, which resulted between CrO3Cl− and both the silanol group (Si/ZrO2 and Si–O–Zr) and siloxane (Si–O–Si) groups of the sol-gel/PUFs [92]. All these processes most likely participated in the CrO3Cl− absorption. On the other hand, “surface adsorption” on/in the sol-gel/PUFs membrane [56] may also participate in the CrO3Cl− sorption. Thus, the overall sorption mechanism of chromium (VI) by the established sol-gel/PQ+.Cl−/PUFs is most likely expressed by the following equation [62, 78]:
| (18) |
where Cr and Caq are the equilibrium concentration of chromium (VI) onto the solid sorbent and the remaining concentration in the aqueous solution, respectively. The parameters Cabs and Cads are the equilibrium concentration of chromium (VI) onto the sorbent as an absorbed and adsorbed species, respectively. S and KL are the saturation values for the Langmuir adsorption.
These result in addition to the resilience characteristics of the PUFs, sol-gel coating technology may also enrich the porous structure of the PUFs, ensuing an improvement in the extraction efficiency of the CrO3Cl− anion via polycondensation of the residual hydroxyl Si–OH and Zr–OH into hydrophobic Si–O–Si and Si–O–Zr groups of the sol-gel as reported [92]. These results suggest the use of the established sol-gel-treated PUFs microextractor in column mode for complete removal and recovery of oxychromium (VI) species in environmental water samples.
3.7. Interference Study
The developed extractor towards CrO3Cl− (10.0 µg·mL−1) sorption from the water was studied via the batch mode of separation in the presence of a series of coexciting inorganic ions, which are commonly present in environmental water samples. The tolerance limit (w/v) of less than ±5% change in the extraction percentage of CrO3Cl− was considered free of interfering species. The impact of the cations NH4+, Na+, K+, Cr3+, Cu2+, Zn2+, Al3+, and Ni2+, as well as the anions NO3−, CH3COO−, F−, Cl−, and Br− at a high mass excess (50-fold) was tested. The data revealed a nonsignificant interference of these ions on chromium (VI) uptake. These results added further support to the utility of the developed extractor towards the removal of chromium (VI) from environmental water samples.
4. Analytical Applications
4.1. Chromium (VI) Retention and Recovery (R) by the Sorbent-Packed Minicolumn Mode
The proposed sol-gel/PQ+.Cl−/PUFs were successfully implemented for complete enrichment and recovery of standard concentrations (0.01–1.0 µg·mL−1) of chromium (VI) spiked into deionized water and tap water samples as described above. The results revealed complete sorption of chromium (VI), as indicated from the absorbance of the total effluent solution. The sorbed chromium (VI) species on the sol-gel/PQ+.Cl−/PUFs were then recovered quantitatively from the sorbent-packed column. The analytical results for chromium (VI) spiked into deionized water and tap water are summarized in Table 2. These results confirm the analytical utility of the established solid sorbent for the complete removal and recovery of chromium (VI) species from water samples at a reasonable flow rate.
Table 2.
Recovery percentage (%) of chromium (VI) spiked into deionized and tap water samples by the established sorbent-packed minicolumn at a 10.0 mL/min flow rate.
| Matrix | Chromium (VI) added (µg/mL) | Chromium (VI) found (µg/mL) | Recovery (%) |
|---|---|---|---|
| Deionized water | 0.01 | 0.01 ± 0.00004 | 100 ± 0.4 |
| 0.10 | 0.1 ± 0.0005 | 100 ± 0.5 | |
| 1.0 | 0.95 ± 0.004 | 95 ± 0.4 | |
|
| |||
| Tap water | 0.01 | 0.01 ± 0.00008 | 100 ± 0.8 |
| 0.10 | 0.1 ± 0.0006 | 100 ± 0.6 | |
| 1.0 | 0.96 ± 0.004 | 96 ± 0.4 | |
4.2. Chromium (VI) Retention and Recovery (R) by the Established Sorbent Packed Pulse Column
In this experiment, a series of prefiltered tap water samples spiked with known concentrations (0.01–1.0 µg/mL) of chromium (VI) at the optimized parameters of extraction were extracted by the sol-gel/PQ+.Cl−/PUFs packed medical syringe (0.2 ± 0.01 g) as a pulse column, as described above. By 25–30 pulses, complete retention (98 ± 2.6%) of the chromium (VI) species was achieved, as indicated by the absorbance of the test acidic aqueous solution after extraction at λmax against the reagent blank (Table 3).
Table 3.
Recovery percentage of chromium (VI) spiked into tap water samples by the established sorbent-packed pulse column (0.2 ± 0.001 g) after 25–30 pulses.
| Chromium (VI) added (µg/mL) | Chromium (VI) found (µg/mL) | Recovery (%) |
|---|---|---|
| 0.01 | 0.01 ± 0.00008 | 100 ± 0.8 |
| 0.05 | 0.05 ± 0.00035 | 100 ± 0.7 |
| 0.50 | 0.5 ± 0.003 | 100 ± 0.6 |
| 1.0 | 0.95 ± 0.006 | 95 ± 0.6 |
The extraction efficiency (En,op) in the open arrangement is related to the distribution ratio (KD), the number of pulses (n), and the maximum volume concentration (p) (p = W0/VP), where W0 is the total sample volume and VP is the compressed foam volume as expressed by the following equation:
| (19) |
The retained chromium (VI) species onto the sorbent-packed pulse was then completely recovered from the sorbent packed pulse column using NaOH (20 mL, 0.5 M) for 25–30 pulses, as indicated from the absorbance of the acidic test solution at 355 nm. The results are summarized in Table 3 and confirm the good performance of the established d–µ SPME towards complete extraction and recovery of chromium (VI) from environmental water samples. The data added further support the utility of the solid sorbent for the complete removal and recovery of chromium (VI) species from water samples by a pulse-packed column.
5. Conclusion and Future Perspective
In summary, the current study reports a facile preparation of sol-gel-functionalized polyurethane foams as an effective dispersive solid-phase microextractor for complete removal and subsequent determination of chromium (VI). The treated foams can serve as an advanced dispersive solid-phase microextractor due to their attractive resilience, flexibility, adjustable surface characteristics, and great capability to extract a series of inorganic and organic species from an aqueous solution. Nanosized sol-gel provides high surface area, H-bonding, and specific affinity for analyte retention on a solid-phase extractor such as PUFs. The membrane-like structure of PUFs solid sorbent facilitates its analytical utility as an effective and resilient solid platform sorbent material for the complete sorption of chromium (VI) species from an aqueous media. In future studies, the established extractor can be used for the complete extraction of chromium (III) and chromium (VI) after the complete oxidation of the former species with H2O2 in an alcoholic KOH solution.
Acknowledgments
This project was funded by the Deanship of Scientific Research (DSR) at King Abdulaziz University, Jeddah, Saudi Arabia, under grant no. (G: 801-247-1441). The author, therefore, acknowledges DSR for technical and financial support.
Data Availability
Electronic supporting information (ESI) will be available online.
Conflicts of Interest
The author declares there are no conflicts of interest.
Supplementary Materials
Table 1 shows the operational conditions of ICP-OES of chromium determination, while scheme 1 and the next four figures show the preparation and characterization of the sol-gel. Figures 7 to 11 show the relation between chromium retention and different parameters such as time, temperature, and concentration.
References
- 1.Martin S., Griswold W. Human health effects of heavy metals. Environmental Science and Technology Briefs for Citizens . 2009;15:1–6. [Google Scholar]
- 2.Bielicka A., Bojanowska I., Wiśniewski A. Two Faces of Chromium- Pollutant and Bioelement. Polish Journal of Environmental Studies . 2005;14:5–10. [Google Scholar]
- 3.Priego-Capote F., Luque de Castro M. Speciation of chromium by in-capillary derivatization and electrophoretically mediated microanalysis. Journal of Chromatography A . 2006;1113(1-2):244–250. doi: 10.1016/j.chroma.2006.01.122. [DOI] [PubMed] [Google Scholar]
- 4.US-EPA. National Priorities List (NPL) Sites with Fiscal Year 1982–2003, Records of Decision (RODs). Office of Emergency and Remedial Response . Washington, DC, USA: US-EPA; 2003. [Google Scholar]
- 5.Manahan S. Environmental Chemistry . Boca Raton, FL, USA: CRC press; 2017. [Google Scholar]
- 6.Reeve R. N. Introduction to Environmental Analysis . Hoboken, NY, USA: John Wiley and Sons; 2002. [Google Scholar]
- 7.US-EPA. Integrated Risk Information System (IRIS) on Chromium VI . Washington, DC, USA: US, Government Printing Office; 1999. [Google Scholar]
- 8.CCME. Canadian soil Quality Guidelines for the Protection of Environmental and Human Health . Winnipeg, Canada: Council of Ministers of the Environment; 2015. [Google Scholar]
- 9.Gettens R. J., Stout G. L. Painting materials: a short encyclopaedia. Courier Corporation . 1966;12 [Google Scholar]
- 10.Gallios G. P., Vaclavikova M. Removal of chromium (VI) from water streams: a thermodynamic study. Environmental Chemistry Letters . 2008;6(4):235–240. doi: 10.1007/s10311-007-0128-8. [DOI] [Google Scholar]
- 11.Ashraf A., Bibi I., Niazi N. K., et al. Chromium(VI) sorption efficiency of acid-activated banana peel over organo-montmorillonite in aqueous solutions. International Journal of Phytoremediation . 2017;19(7):605–613. doi: 10.1080/15226514.2016.1256372. [DOI] [PubMed] [Google Scholar]
- 12.Portier C. J. Toxicological Profile for Chromium . Washington, DC, USA: U.S. Department of Health and Human Services; 2012. [Google Scholar]
- 13.Ahmad W., Bashammakh A., Al-Sibaai A., Alwael H., El-Shahawi M. Trace determination of Cr(III) and Cr(VI) species in water samples via dispersive liquid-liquid microextraction and microvolume UV–Vis spectrometry. Thermodynamics, speciation study. Journal of Molecular Liquids . 2016;224:1242–1248. doi: 10.1016/j.molliq.2016.10.106. [DOI] [Google Scholar]
- 14.Jie N., Zhang Q., Yang J., Huang X. Determination of chromium in waste-water and cast iron samples by fluorescence quenching of rhodamine 6G. Talanta . 1998;46(1):215–219. doi: 10.1016/s0039-9140(97)00261-0. [DOI] [PubMed] [Google Scholar]
- 15.Jia-you G. A. O. Determination of chromium (VI) with rhodamine 6G by fluorescence quenching method. Metallurgical Analysis . 2005;1:p. 009. [Google Scholar]
- 16.Mahesvari V., Balasabramanian N. Spectrophotometric determination of chromium based on ion-pair formation. Chemia Analityczna . 1996;41:569–576. [Google Scholar]
- 17.Alexovič M., Balogh I. S., Škrlíková J., Andruch V. A. A dispersive liquid–liquid microextraction procedure for UV-Vis spectrophotometric determination of chromium (VI) in water samples. Analytical Methods . 2012;4(5):1410–1414. doi: 10.1039/c2ay25133g. [DOI] [Google Scholar]
- 18.Dehghani Mohammad Abadi M., Chamsaz M., Arbab-Zavar M. H., Shemirani F. Supramolecular dispersive liquid–liquid microextraction based solidification of floating organic drops for speciation and spectrophotometric determination of chromium in real samples. Analytical Methods . 2013;5(12):2971–2977. doi: 10.1039/c3ay00036b. [DOI] [Google Scholar]
- 19.Telepčáková M., Andruch V., Balogh I. S. Indirect extraction-spectrophotometric determination of chromium. Chemical Papers . 2005;59:109–112. [Google Scholar]
- 20.Kong Z., Wei J., Guo H., Wang H., Li Y., Liu N. The determination of trace amount of Cr (VI) and As (V) in aqueous solution by UV–Vis spectrophotometer assisted with preconcentration on a quaternary ammonium ion exchange fiber-in-tube. Desalination and Water Treatment . 2015;53(2):382–389. doi: 10.1080/19443994.2013.841103. [DOI] [Google Scholar]
- 21.Alshammari M. S., Ahmed I. M., Alsharari J. S., et al. Adsorption of Cr(VI) using α-Fe2O3 coated hydroxy magnesium silicate (HMS): isotherm, thermodynamic and kinetic study. International Journal of Environmental Analytical Chemistry . 2021;103(10):2223–2239. doi: 10.1080/03067319.2021.1890061. [DOI] [Google Scholar]
- 22.El-Shahawi M. S., Hassan S. S. M., Othman A. M., Zyada M. A., El-Sonbati M. A. Chemical speciation of chromium (III, VI) employing extractive spectrophotometry and tetraphenylarsonium chloride or tetraphenylphosphonium bromide as ion-pair reagent. Analytica Chimica Acta . 2005;534:319–326. [Google Scholar]
- 23.Hemmatkhah P., Bidari A., Jafarvand S., Milani Hosseini M. R., Assadi Y. Speciation of chromium in water samples using dispersive liquid–liquid microextraction and flame atomic absorption spectrometry. Microchimica Acta . 2009;166(1-2):69–75. doi: 10.1007/s00604-009-0167-x. [DOI] [Google Scholar]
- 24.Pereira Lara P. C., Nicácio Silveira J., Borges Neto W., Beinner M. A., da Silva J. B. Use of multivariate optimization to develop a method for direct chromium determination in breast milk by GF-AAS using aqueous calibration. Journal of Chemical and Pharmaceutical Research . 2015;7:1900–1906. [Google Scholar]
- 25.Borges A. R., François L. L., Becker E. M., Vale M. G. R, Welz B. Method development for the determination of chromium and thallium in fertilizer samples using graphite furnace atomic absorption spectrometry and direct solid sample analysis. Microchemical Journal . 2015;119:169–175. doi: 10.1016/j.microc.2014.11.007. [DOI] [Google Scholar]
- 26.Hong Z. T., Jin L., Zhi P. Q., Ying Z., Jie L. Direct determination of trace chromium in milk powder by graphite furnace atomic absorption spectrometry with suspension sampling. Journal of Food Safety and Quality . 2015;6:1478–1482. [Google Scholar]
- 27.Saputro S., Yoshimura K., Matsuoka S., et al. Speciation of dissolved chromium and the mechanisms controlling its concentration in natural water. Chemical Geology . 2014;364:33–41. doi: 10.1016/j.chemgeo.2013.11.024. [DOI] [Google Scholar]
- 28.Petrucci F., Senofonte O. Determination of Cr (VI) in cosmetic products using ion chromatography with dynamic reaction cell inductively coupled plasma mass spectrometry (DRC-ICP-MS) Analytical Methods . 2015;7(12):5269–5274. doi: 10.1039/c4ay03042g. [DOI] [Google Scholar]
- 29.Goncalves D. A., Gu J., dos Santos M. C., Jones B. T., Donati G. L. Direct determination of chromium in empty medicine capsules by tungsten coil atomic emission spectrometry. Journal of Analytical Atomic Spectrometry . 2015;30(6):1395–1399. doi: 10.1039/c5ja00100e. [DOI] [Google Scholar]
- 30.Vidal L., Silva S. G., Canals A., Nóbrega J. A. Tungsten coil atomic emission spectrometry combined with dispersive liquid–liquid microextraction: A synergistic association for chromium determination in water samples. Talanta . 2016;148:602–608. doi: 10.1016/j.talanta.2015.04.023. [DOI] [PubMed] [Google Scholar]
- 31.Bahadir Z., Bulut V. N., Hidalgo M., Soylak M., Margui E. Cr Speciation in water samples by dispersive liquid-liquid microextraction combined with total reflection X-ray fluorescence spectrometry. Spectrochimica Acta Part B: Atomic Spectroscopy . 2016;115:46–51. doi: 10.1016/j.sab.2015.11.001. [DOI] [Google Scholar]
- 32.Wang L. L., Wang J. Q., Zheng Z. X., Xiao P. Cloud point extraction combined with high performance liquid chromatography for speciation of chromium (III) and chromium (VI) in environmental sediment samples. Journal of Hazardous Materials . 2010;177(1–3):114–118. doi: 10.1016/j.jhazmat.2009.12.003. [DOI] [PubMed] [Google Scholar]
- 33.Tang A. N., Jiang D. Q., Jiang Y., Wang S. W., Yan X. P. Cloud point extraction for high-performance liquid chromatographic speciation of Cr(III) and Cr(VI) in aqueous solutions. Journal of Chromatography A . 2004;1036(2):183–188. doi: 10.1016/j.chroma.2004.02.065. [DOI] [PubMed] [Google Scholar]
- 34.Sun J., Yang Z., Lee H., Wang L. Simultaneous speciation and determination of arsenic, chromium and cadmium in water samples by high performance liquid chromatography with inductively coupled plasma mass spectrometry. Analytical Methods . 2015;7(6):2653–2658. doi: 10.1039/c4ay02813a. [DOI] [Google Scholar]
- 35.Catalani S., Fostinelli J., Gilberti M. E., Apostoli P. Application of a metal free high performance liquid chromatography with inductively coupled plasma mass spectrometry (HPLC–ICP-MS) for the determination of chromium species in drinking and tap water. International Journal of Mass Spectrometry . 2015;387:31–37. doi: 10.1016/j.ijms.2015.06.015. [DOI] [Google Scholar]
- 36.Li P., Li L. M., Xia J., et al. Determination of hexavalent chromium in traditional chinese medicines by high performance liquid chromatography with inductively coupled plasma mass spectrometry. Journal of Separation Science . 2015;38(23):4043–4047. doi: 10.1002/jssc.201500814. [DOI] [PubMed] [Google Scholar]
- 37.Vacchina V., de la Calle I., Séby F. Cr (VI) speciation in foods by HPLC-ICP-MS: investigation of Cr (VI)/food interactions by size exclusion and Cr (VI) determination and stability by ion-exchange on-line separations. Analytical and Bioanalytical Chemistry . 2015;407(13):3831–3839. doi: 10.1007/s00216-015-8616-3. [DOI] [PubMed] [Google Scholar]
- 38.Rezaee M., Assadi Y., Milani Hosseini M. R., Aghaee E., Ahmadi F., Berijani S. Determination of organic compounds in water using dispersive liquid–liquid microextraction. Journal of chromatography A . 2006;1116(1-2):1–9. doi: 10.1016/j.chroma.2006.03.007. [DOI] [PubMed] [Google Scholar]
- 39.Kocúrová L., Balogh I. S., Šandrejová J., Andruch V. Recent advances in dispersive liquid–liquid microextraction using organic solvents lighter than water. A review. Microchemical Journal . 2012;102:11–17. doi: 10.1016/j.microc.2011.12.002. [DOI] [Google Scholar]
- 40.Aldawsari A. M., Alsohaimi I. H., Hassan H. M. A., Al-Abduly A. J., Hassan I., Saleh E. A. M. Multiuse silicon hybrid polyurea-based polymer for highly effective removal of heavy metal ions from aqueous solution. International journal of Environmental Science and Technology . 2022;19(4):2925–2938. doi: 10.1007/s13762-021-03355-6. [DOI] [Google Scholar]
- 41.Chequer F. M. D., Angeli J. P. F., Ferraz E. R. A., et al. The azo dyes disperse red 1 and disperse orange 1 increase the micronuclei frequencies in human lymphocytes and in HepG2 cells. Mutation Research/Genetic Toxicology and Environmental Mutagenesis . 2009;676(1-2):83–86. doi: 10.1016/j.mrgentox.2009.04.004. [DOI] [PubMed] [Google Scholar]
- 42.El-Shahawi M. S. Retention and separation of some organic water pollutants with unloaded and tri-n-octylamine loaded polyester-based polyurethane foams. Talanta . 1994;41(9):1481–1488. doi: 10.1016/0039-9140(94)e0039-t. [DOI] [PubMed] [Google Scholar]
- 43.Sharma R., N Chavan S., Lee H. Fluorescent imidazolium hydrogels for tracing and recovering platinum with highest purity from spent auto catalyst. Sensors and Actuators B: Chemical . 2023;396 doi: 10.1016/j.snb.2023.134625.134625 [DOI] [Google Scholar]
- 44.Hameed B. H. Spent tea leaves: a new non-conventional and low-cost adsorbent for removal of basic dye from aqueous solutions. Journal of Hazardous Materials . 2009;161(2-3):753–759. doi: 10.1016/j.jhazmat.2008.04.019. [DOI] [PubMed] [Google Scholar]
- 45.Bowen H. J. M. Absorption by polyurethane foams: new method of separation. Journal of the Chemical Society A: Inorganic, Physical, Theoretical . 1970;7:1082–1085. doi: 10.1039/j19700001082. [DOI] [Google Scholar]
- 46.de Jesus da Silveira Neta J., Costa Moreira G., da Silva C. J., Reis C., Reis E. L. Use of polyurethane foams for the removal of the Direct Red 80 and Reactive Blue 21dyes in aqueous medium. Desalination . 2011;281:55–60. doi: 10.1016/j.desal.2011.07.041. [DOI] [Google Scholar]
- 47.Ruixia L., Jinlong G., Hongxiao T. Adsorption of fluoride, phosphate, and arsenate ions on a new type of ion exchange fiber. Journal of Colloid and Interface Science . 2002;248(2):268–274. doi: 10.1006/jcis.2002.8260. [DOI] [PubMed] [Google Scholar]
- 48.Cao F., Jaunat J., Sturchio N., et al. Worldwide occurrence and origin of perchlorate ion in waters: A review. Science of the Total Environment . 2019;661:737–749. doi: 10.1016/j.scitotenv.2019.01.107. [DOI] [PubMed] [Google Scholar]
- 49.Xie Y., Ren L., Zhu X., Gou X., Chen S. Physical and chemical treatments for removal of perchlorate from water–a review. Process Safety and Environmental Protection . 2018;116:180–198. doi: 10.1016/j.psep.2018.02.009. [DOI] [Google Scholar]
- 50.Alwael H., Alsulami A. N., Abduljabbar T. N., Oubaha M., El-Shahawi M. S. Innovative Sol-gel functionalized polyurethane foam for sustainable water purification and analytical advances. Frontiers in Chemistry . 2024;12 doi: 10.3389/fchem.2024.1324426.1324426 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Alsohaimi I. H., Alhumaimess M. S., Alqadami A. A., et al. Chitosan-carboxylic acid grafted multifunctional magnetic nanocomposite as a novel adsorbent for effective removal of methylene blue dye from aqueous environment. Chemical Engineering Science . 2023;280 doi: 10.1016/j.ces.2023.119017.119017 [DOI] [Google Scholar]
- 52.Hassan H. M. A, El-Aassar M. R., El-Hashemy M. A., et al. Sulfanilic acid-functionalized magnetic GO as a robust adsorbent for the efficient adsorption of methylene blue from aqueous solution. Journal of Molecular Liquids . 2022;361 doi: 10.1016/j.molliq.2022.119603.119603 [DOI] [Google Scholar]
- 53.Oubaha M., Gorin A., McDonagh C., Duffy B., Copperwhite R. Development of a multianalyte optical sol–gel biosensor for medical diagnostic. Sensors and Actuators B: Chemical . 2015;221:96–103. doi: 10.1016/j.snb.2015.06.012. [DOI] [Google Scholar]
- 54.Cullen M., Morshed M., O’Sullivan M., MacHugh E., Duffy B., Oubaha M. Correlation between the structure and the anticorrosion barrier properties of hybrid sol–gel coatings: application to the protection of AA2024-T3 alloys. Journal of Sol-Gel Science and Technology . 2017;82(3):801–816. doi: 10.1007/s10971-017-4349-4. [DOI] [Google Scholar]
- 55.Nakamoto N. Infrared spectra of inorganic coordination compounds . 4th. New York, NY, USA: Wiley, Inter science; 1986. [Google Scholar]
- 56.Gunashekar S., Abu-Zahra N. Synthesis of functionalized polyurethane foam using BES chain extender for lead ion removal from aqueous solutions. Journal of Cellular Plastics . 2015;51(5–6):453–470. doi: 10.1177/0021955x14559255. [DOI] [Google Scholar]
- 57.Aldawsari A. M., Alsohaimi I. H., Hassan H. M. A., Abdalla Z. E. A., Hassan I., Berber M. R. Tailoring an efficient nanocomposite of activated carbon-layered double hydroxide for elimination of water-soluble dyes. Journal of Alloys and Compounds . 2021;857 doi: 10.1016/j.jallcom.2020.157551.157551 [DOI] [Google Scholar]
- 58.Zhang X., Ju H., Wang J. Electrochemical Sensors, Biosensors and their Biomedical Applications . Cambridge, UK: Academic Press; 2008. [Google Scholar]
- 59.Sharma R., Haldar U., Turabee M. H., Lee H. I. Recyclable macromolecular thermogels for Hg(II) detection and separation via sol-gel transition in complex aqueous environments. Journal of Hazardous Materials . 2021;410 doi: 10.1016/j.jhazmat.2020.124625.124625 [DOI] [PubMed] [Google Scholar]
- 60.Sharma R., Suhendra N. F., Jung S., Lee H. Gold recovery at ultra-high purity from electronic waste using selective polymeric film. Chemical Engineering Journal . 2023;451(1) doi: 10.1016/j.cej.2022.138506.138506 [DOI] [Google Scholar]
- 61.Al-Bagawi A. H., Mohammed G. I., Ahmad W., Alwael H., Moustafa Y. M., El-Shahawi M. S. Chromatographic separation total determination and chemical speciation of mercury in environmental water samples using 4-(2-thiazolylazo) resorcinol based polyurethane foam sorbent packed column. Chemometric and Data Analysis in Chromatography . 2018;23 [Google Scholar]
- 62.El-Shahawi M. S., Alwael H. A novel platform based on gold nanoparticles chemically impregnated polyurethane foam sorbent coupled ion chromatography for selective separation and trace determination of phosphate ions in water. Microchemical Journal . 2019;249:00987–100997. [Google Scholar]
- 63.Braun T., Navratil J. D., Farag A. B. Polyurethane foam sorbents in separation science . 1st. Boca Raton, FL, USA: CRC Press; 1985. [Google Scholar]
- 64.Vogel A. I., Bassett R. C. D., Jeffrey G. H. A Text-book of Quantitative Inorganic Analysis: Including Elementary Instrumental Analysis . 3rd. England, UK: Longman; 1952. [Google Scholar]
- 65.Gupta V. K., Gupta M., Sharma S. Process development for the removal of lead and chromium from aqueous solutions using red mud - an aluminium industry waste. Water Research . 2001;35(5):1125–1134. doi: 10.1016/s0043-1354(00)00389-4. [DOI] [PubMed] [Google Scholar]
- 66.Tuzen M., Soylak M. Multiwalled carbon nanotubes for speciation of chromium in environmental samples. Journal of Hazardous Materials . 2007;147(1–2):219–225. doi: 10.1016/j.jhazmat.2006.12.069. [DOI] [PubMed] [Google Scholar]
- 67.Mohammed T., Ibrahim R., Naji A. Experimental investigation and thermodynamic study of heavy metal removal from industrial wastewater using pomegranate peel. MATEC Web of Conferences . 2018;162:p. 05007. doi: 10.1051/matecconf/201816205007.05007 [DOI] [Google Scholar]
- 68.Aydın Y. A., Aksoy N. D. Adsorption of chromium on chitosan: optimization, kinetics and thermodynamics. Chemical Engineering Journal . 2009;151(1–3):188–194. doi: 10.1016/j.cej.2009.02.010. [DOI] [Google Scholar]
- 69.Bhattacharya A. K., Naiya T. K., Mandal S. N., Das S. K. Adsorption, kinetics and equilibrium studies on removal of Cr(VI) from aqueous solutions using different low-cost adsorbents. Chemical Engineering Journal . 2008;137:529–541. [Google Scholar]
- 70.Ntuli T. D., Pakade V. E. Hexavalent chromium removal by polyacrylic acid-grafted macadamia nutshell powder through adsorption–reduction mechanism: adsorption isotherms, kinetics and thermodynamics. Chemical Engineering Communications . 2020;207(3):279–294. doi: 10.1080/00986445.2019.1581619. [DOI] [Google Scholar]
- 71.Parlayici Ş., Pehlivan E. Comparative study of Cr(VI) removal by bio-waste adsorbents: equilibrium, kinetics, and thermodynamic. Journal of Analytical Science and Technology . 2019;10(1) doi: 10.1186/s40543-019-0175-3. [DOI] [Google Scholar]
- 72.Alouache A., Selatnia A., Sayah H. E., Khodja M., Moussous S., Daoud N. Biosorption of hexavalent chromium and congo red dye onto Pleurotus mutilus biomass in aqueous solutions. International Journal of Environmental Science and Technology . 2022;19(4):2477–2492. doi: 10.1007/s13762-021-03313-2. [DOI] [Google Scholar]
- 73.MacHugh E., Cullen M., Kaworek A., Duffy B., Oubaha M. The effect of curing and zirconium content on the wettability and structure of a silicate hybrid sol-gel material. Journal of Non-Crystalline Solids . 2019;525 doi: 10.1016/j.jnoncrysol.2019.119658.119658 [DOI] [Google Scholar]
- 74.Zhang J. L., Kong J. Z., Li A. D., et al. Synthesis and characterization of FePt nanoparticles and FePt nanoparticle/SiO2-matrix composite films. Journal of Sol-Gel Science and Technology . 2010;64(2):269–275. doi: 10.1007/s10971-010-2373-8. [DOI] [Google Scholar]
- 75.El-Shahawi M. S., Al-Saidi H. M., Bashammakh A. S., Al-Sibaai A. A., Abdelfadeel M. A. Spectrofluorometric determination and chemical speciation of trace concentrations of chromium (III and VI) species in water using the ion pairing reagent tetraphenyl- phosphonium bromide. Talanta . 2011;84(1):175–179. doi: 10.1016/j.talanta.2010.12.039. [DOI] [PubMed] [Google Scholar]
- 76.Tandon R. K., Crisp P. T., Ellis J., Baker R. S. Effect of pH on chromium (VI) species in solution. Talanta . 1984;31(3):227–228. doi: 10.1016/0039-9140(84)80059-4. [DOI] [PubMed] [Google Scholar]
- 77.Balogh I. S., Maga I. M., Hargitai-Toth A., Andruch V. Spectrophotometric study of the complexation and extraction of chromium (VI) with cyanine dyes. Talanta . 2000;53(3):543–549. doi: 10.1016/s0039-9140(00)00525-7. [DOI] [PubMed] [Google Scholar]
- 78.Bashammakh A. The retention profile of phosphate ions in aqueous media onto ion pairing immobilized polyurethane foam: Kinetics, sorption and chromatographic separation. Journal of Molecular Liquids . 2016;220:426–431. doi: 10.1016/j.molliq.2016.04.105. [DOI] [Google Scholar]
- 79.Othmani A., Magdouli S., Senthil Kumar P., Kapoor A., Chellam P. V., Gokkus O. Agricultural waste materials for adsorptive removal of phenols, chromium (VI) and cadmium (II) from wastewater: A review. Environmental Research . 2022;204 doi: 10.1016/j.envres.2021.111916.111916 [DOI] [PubMed] [Google Scholar]
- 80.Kerur S. S., Bandekar S., Hanagadakar M. S., Nandi S. S., Ratnamala G. M., Hegde P. G. Removal of hexavalent chromium-industry treated water and wastewater: A review. Materials Today: Proceedings . 2021;42:1112–1121. doi: 10.1016/j.matpr.2020.12.492. [DOI] [Google Scholar]
- 81.Mohammad S. G., Ahmed S. M. Preparation of environmentally friendly activated carbon for removal of pesticide from aqueous media. International Journal of Industrial Chemistry . 2017;8(2):121–132. doi: 10.1007/s40090-017-0115-2. [DOI] [Google Scholar]
- 82.Jalili V., Barkhordari A., Ghiasvand A. A comprehensive look at solid-phase microextraction technique: A review of reviews. Microchemical Journal . 2020;152 doi: 10.1016/j.microc.2019.104319.104319 [DOI] [Google Scholar]
- 83.Murtada K. Trends in nanomaterial-based solid-phase microextraction with a focus on environmental applications-A review. Trends in Environmental Analytical Chemistry . 2020;25 doi: 10.1016/j.teac.2019.e00077. [DOI] [Google Scholar]
- 84.Ima E. C., Adebayo M. A., Machado F. M. Kinetic and Equilibrium Models of Adsorption. In: Bergmann C., Machado F., editors. Carbon Nanomaterials as Adsorbents for Environmental and Biological Applications. Carbon Nanostructures . Berlin, Germany: Springer; 2015. [Google Scholar]
- 85.Weber W. J. Evolution of a Technology. Journal of Environmental Engineering . 1984;110(5):899–917. doi: 10.1061/(asce)0733-9372(1984)110:5(899). [DOI] [Google Scholar]
- 86.Lagergren K. S. About the theory of so-called adsorption of soluble substances. Kungliga Svenska Vetenskapsakademiens Handlingar . 1898;24:1–39. [Google Scholar]
- 87.Ho Y. S., McKay G. Pseudo-second order model for sorption processes. Process Biochemistry . 1999;34(5):451–465. doi: 10.1016/s0032-9592(98)00112-5. [DOI] [Google Scholar]
- 88.McLintock I. S. The Elovich equation in chemisorption kinetics. Nature . 1967;216(5121):1204–1205. doi: 10.1038/2161204a0. [DOI] [Google Scholar]
- 89.Vijayaraghavan K., Palanivelu K., Velan M. Biosorption of copper(II) and cobalt(II) from aqueous solutions by crab shell particles. Bioresource Technology . 2006;97(12):1411–1419. doi: 10.1016/j.biortech.2005.07.001. [DOI] [PubMed] [Google Scholar]
- 90.Ingrand V., Guinamant J. L., Bruchet A., et al. Determination of bromate in drinking water: development of laboratory and field methods. TrAC Trends in Analytical Chemistry . 2002;21(1):1–12. doi: 10.1016/s0165-9936(01)00113-3. [DOI] [Google Scholar]
- 91.Huang X., Mao X. Y., Bu H. T., Yu X. Y., Jiang G. B., Zeng M. H. Chemical modification of chitosan by tetraethylenepentamine and adsorption study for anionic dye removal. Carbohydrate Research . 2011;346(10):1232–1240. doi: 10.1016/j.carres.2011.04.012. [DOI] [PubMed] [Google Scholar]
- 92.Olu-owolabi B. I., Diagboya P. N., Adebowale K. O. Evaluation of pyrene sorption -desorption on tropical soils. Journal of Environmental Management . 2014;137:1–9. doi: 10.1016/j.jenvman.2014.01.048. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Table 1 shows the operational conditions of ICP-OES of chromium determination, while scheme 1 and the next four figures show the preparation and characterization of the sol-gel. Figures 7 to 11 show the relation between chromium retention and different parameters such as time, temperature, and concentration.
Data Availability Statement
Electronic supporting information (ESI) will be available online.
