Next Article in Journal
Role of Peripheral Coordination Boron in Electrocatalytic Nitrogen Reduction over N-Doped Graphene-Supported Single-Atom Catalysts
Next Article in Special Issue
β-CD-Induced Precipitation of Eriochrome Black T Recovered via CTAB-Assisted Foam Fractionation for Adsorption of Trace Cu(II)
Previous Article in Journal
Analysis of Aminoglycoside Antibiotics: A Challenge in Food Control
Previous Article in Special Issue
Simple Co-Precipitation of Iron Minerals for the Removal of Phenylarsonic Acid: Insights into the Adsorption Performance and Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pb(II) Uptake from Polluted Irrigation Water Using Anatase TiO2 Nanoadsorbent

by
Miguel A. Vasquez-Caballero
1,
Yamerson Canchanya-Huaman
1,
Angie F. Mayta-Armas
1,
Jemina Pomalaya-Velasco
1,
Noemi-Raquel Checca-Huaman
2,
Yéssica Bendezú-Roca
1 and
Juan A. Ramos-Guivar
3,*
1
Laboratorio de No Metálicos, Facultad de Ingeniería Química, Universidad Nacional del Centro del Perú (UNCP), Av. Mariscal Ramón Castilla Nº 3909, El Tambo, Huancayo 12000, Peru
2
Centro Brasileiro de Pesquisas Físicas, Rio de Janeiro 22290-180, RJ, Brazil
3
Grupo de Investigación de Nanotecnología Aplicada para Biorremediación Ambiental, Energía, Biomedicina y Agricultura (NANOTECH), Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Av. Venezuela Cdra 34 S/N, Ciudad Universitaria, Lima 15081, Peru
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(12), 4596; https://doi.org/10.3390/molecules28124596
Submission received: 2 May 2023 / Revised: 20 May 2023 / Accepted: 22 May 2023 / Published: 7 June 2023
(This article belongs to the Special Issue Adsorbents in Treatment of Pollutants)

Abstract

:
The adsorption characteristics of titanium dioxide nanoparticles (nano-TiO2) for the removal of Pb(II) from irrigation water were investigated in this work. To accomplish this, several adsorption factors, such as contact time and pH, were tested to assess adsorption efficiencies and mechanisms. Before and after the adsorption experiments, commercial nano-TiO2 was studied using X-ray diffraction (XRD), scanning and transmission electron microscopy (SEM and TEM), energy dispersive spectroscopy (EDS), and X-ray photoelectron spectroscopy (XPS). The outcomes showed that anatase nano-TiO2 was remarkably efficient in cleaning Pb(II) from water, with a removal efficiency of more than 99% after only one hour of contact time at a pH of 6.5. Adsorption isotherms and kinetic adsorption data matched the Langmuir and Sips models quite well, showing that the adsorption process occurred at homogenous sites on the surface of nano-TiO2 by forming a Pb(II) adsorbate monolayer. The XRD and TEM analysis of nano-TiO2 following the adsorption procedure revealed a non-affected single phase (anatase) with crystallite sizes of 9.9 nm and particle sizes of 22.46 nm, respectively. According to the XPS data and analyzed adsorption data, Pb ions accumulated on the surface of nano-TiO2 through a three-step mechanism involving ion exchange and hydrogen bonding mechanisms. Overall, the findings indicate that nano-TiO2 has the potential to be used as an effective and long-lasting mesoporous adsorbent in the treatment and cleaning of Pb(II) from water bodies.

1. Introduction

Pollution of the environment with heavy metals has become a big issue across the world, affecting both industrialized and developing nations [1]. Water pollution and scarcity are the two most serious environmental issues threatening the existence of plants, animals, and people in the twenty-first century. Human activities have disrupted the natural recycling of water and its purifying mechanisms, resulting in a shortage of water for human use [2]. According to the 2018 United Nations Global Water Development Report, industrial and household water consumption will grow far faster than agricultural need, despite agriculture remaining the largest overall consumer [3]. Water is now utilized for agriculture 70% of the time, with irrigation being the most common application. Electroplating, metallurgy, steel, fertilizers, pesticides, textiles, paints, glass, synthetic materials, and mining are among the sectors that discharge heavy-metal-containing wastewater into the environment [4,5]. Even at low concentrations, these pollutants are very hazardous to human health; thus, they must be treated before disposal [6].
Lead, Pb(II), is one of the most damaging heavy metals; its influence on water contamination has sparked international concern [7]. Both natural weathering processes and human-made operations, such as mining, making batteries, and electroplating, can release Pb(II) into the environment [8]. Because of its bioaccumulation, toxicity, and persistence, Pb(II) is an extremely harmful chemical [9,10]. Infants, children under the age of six, the fetus, and pregnant women are the most vulnerable to poor health impacts [11]. The effects on the central nervous system can be especially severe, including cancer, anemia, and damage to the brain and other organ systems [12,13]. The World Health Organization (WHO) suggests keeping the amount of lead in drinking water to a maximum of 0.01 mg L−1 [11].
Several methods have been developed to reduce the levels of Pb(II) and other heavy metals in polluted water, including adsorption, biosorption, chemical precipitation, ion exchange, membrane separation, ion treatment, electrochemical treatment, recovery by evaporation, redox precipitation, and filtration [12,14,15,16]. Adsorption has numerous benefits over the other approaches [17,18]. It is easy to use and there are a wide range of adsorbents that can be reused after the desorption process in various adsorption cycles, allowing for the reasonably priced separation of pollutants [19]. Nowadays, the utilization of nanoparticles (NPs) as an absorbent for metal pollutants is a highly promising option [20]. Nanomaterials offer excellent characteristics for reducing pollutants in water. Due to their small size [21,22] and large surface area, NPs easily disperse in liquids where they can interact with a variety of chemical species [19].
Nanostructures have emerged as a very attractive alternative for water cleanup due to their huge surface-to-volume ratios, which are excellent for surface reactions [23]. Numerous studies have demonstrated that heavy metals, such as Pb(II) and Cd(II), can be effectively removed from aqueous solutions using nano-TiO2. As a result, investigations on the use of nano-TiO2 in wastewater treatment are extremely beneficial to the aquatic environment [24]. Moreover, numerous TiO2-based nanocomposites have been investigated to increase these materials’ adsorption capacity [25,26]. TiO2 polymorphs include anatase, rutile, and brookite, which are the most prevalent. Each of these polymorphs has unique structural, physical, and chemical features that influence their ability for heavy metal adsorption. Anatase, for example, has a bigger surface area and porosity than rutile and brookite, allowing it to adsorb more metals. Yet, under specific pH and temperature circumstances, rutile and brookite are more stable than anatase [27,28]. Despite their remarkable Pb(II) removal properties in simulated aliquots, little literature is available concerning nano-TiO2 adsorption performance in real waters. Certainly, diverse nano-TiO2 polymorphs will exhibit different removal properties in river, lake, and sea samples containing heavy metals, and, more importantly, the conservation of the physicochemical properties of nano-TiO2 will guarantee their environmental sustainability and reuse.
In that sense, the pollution of water bodies by Pb(II) is a current concern—taking in and getting rid of this metal is important and an interesting challenge. This work used nano-TiO2 as an alternative to remove Pb(II) from irrigation water with a corrected starting concentration of 47.54 mg L−1, and its removal effectiveness was evaluated. The water samples were collected from the CIMIRM irrigation canal in Peru’s province of Jauja. Analytic techniques were used to evaluate nano-TiO2 before and after Pb adsorption—this last to ensure their physicochemical and reusability properties. Multiparametric adsorption analysis, including adsorption kinetics, the effect of pH on adsorption, adsorbent dose on adsorption capacity, and the behavior of experimental data in relation to adsorption isotherms, were also investigated. The results indicate that anatase nano-TiO2 is an alternative solution to the Pb(II) water cleaning problem, including modified surface water samples that exceed the permissible levels needed for agricultural irrigation, e.g., [29].

2. Results and Discussions

2.1. Characterization of Nano-TiO2

2.1.1. XRD and Rietveld Analysis

Figure 1a,b shows the refined X-ray diffractograms of nano-TiO2 with the corresponding refinement model after the Pb adsorption process with initial concentrations of 47.54 mg L−1 (TiO2-PbC1) and 1.16 mg L−1 (TiO2-PbC2). Only the pure anatase phase was observed; the refined parameters are summarized in Table 1. Considering anisotropic size broadening, the mean apparent crystallite sizes were calculated using Scherrer’s formula, which may be expressed as a linear combination of spherical harmonics and is provided in the FullProf Suite program [30]. The data corroborated the presence of the tetragonal crystal structure of the anatase phase even after adsorption of Pb(II) for both initial concentrations. The average apparent size of anatase crystallites was 9.9 nm and 8.7 nm, evidencing that there was not much variation in the microstructural parameters (Table 1), even with increasing concentration of Pb. Anatase crystallite size values of 10.1 nm have been reported by Canchanya et al. and of 12.22 nm by Patidar and Jain, with no application in the adsorption process. In comparison with these values, the crystallite size of anatase showed a slight decrease after the Pb(II) adsorption process [31,32].

2.1.2. SEM and EDS Mapping Analysis

Figure 2 and Figure S1 depict the SEM images for two magnifications (a and f), where characteristic granular morphologies of nano-TiO2 powders exposed to two different Pb(II) concentrations are observed. EDS mapping images are observed in Figure 2 and Figure S1b,g and the elemental images for O (c,h), Ti (d,i), and Pb (e,j) are shown. The presence of Pb(II) was corroborated at low and high initial concentrations. Table 2 summarizes the weight percentage for the identified elements after analyzing the EDS spectrum (see Figure S2).

2.1.3. TEM Analysis

Figure 3 and Figure S3 show the effect of adsorption of Pb(II) on nano-TiO2 morphology at different initial adsorbate concentrations of 47.54 mg L−1 and 1.16 mg L−1, with the aim to study the morphology and particle size distribution (PSD). Figure 3a and Figure S3a show that the nano-TiO2 particles presented mostly spherical morphology [33]. In addition, the PSD was plotted from the count of 700 NPs, from which average particle sizes of 22.46 nm (Figure 3b) and 17.1 nm (Figure S3b) were estimated. The adsorption of Pb(II) with a higher initial concentration showed a slight increase in particle size compared to the TiO2 NP size, without application in the adsorption process, of 17.9 nm reported by Canchanya et al. [31]. With a low initial concentration there was no variation in nano-TiO2 particle size. The PSD histogram makes it obvious that the NPs had a wide PSD and that their distribution closely followed a normal curve. The selected area electron diffraction (SAED) patterns are presented in Figure 3d and Figure S3d. The crystalline planes of nano-TiO2 are specified by the Miller indices, which were identified from the pattern rings, which agree with the XRD results (Figure 1). Looking at the images, it can be suggested that the Pb adsorbed on nano-TiO2 formed a polynanocrystalline structure [33,34]. Figure 3c and Figure S3c show the crystal lattice planes of nano-TiO2, for both samples; a crystal lattice plane (110) with a d-spacing of 0.35 nm was estimated. On the other hand, the crystallite domain size distribution gave values of 9.9 nm (Figure 3e) and 8.7 nm (Figure S3e).

2.1.4. XPS Analysis

Figure 4 and Figure 5 show the XPS spectra for low- and high-concentrated Pb adsorbed onto the nano-TiO2 sample—in principle, there was no alteration in the BE of Ti 2p3/2 with respect to both Pb concentrations. Furthermore, by increasing the concentration of the metal (Pb) of the aqueous solution (from 1.16 mg L−1 to 47.54 mg L−1) in the nano-TiO2, no increases in at%. Pb adsorbed were observed (Table 3). This indicates a saturation limit of adsorption of Pb atoms by the TiO2 NPs. Regarding the energy analysis of Ti 2p3/2, this indicated that TiO2 with a rutile or anatase structure (indistinguishable by XPS) [35] went through an oxidation process with respect to its initial structure. This was due to the adsorption process and was confirmed by the BE of O 1s at 531.7 eV, as suggested by Lu et al. [36].

2.2. Adsorption Kinetics

In order to find the time in which maximum removal is obtained, the adsorption kinetic performance was evaluated at pH 5.5, 298 K and a dose of 1 g L−1. This removal percentage was calculated by applying the following Equation (1):
R % = C 0 C f C 0 × 100 %
Figure 6 shows that the percentage removal of Pb(II) on nano-TiO2 was higher than 92% in all cases; with longer time, there were only slight variations. Therefore, the equilibrium time was set at 1 h with 99.18% Pb removal. This time frame was shorter than the value found by Poursani et al., who determined the equilibrium time to be 4 h at an adsorbent dose of 3 g L−1 [39]. Moreover, the study of Recillas et al. reported an equilibrium time of 0.62 h at 1 g L1 [40]. Jyothi et al. achieved 70% Pb removal with a dose of 0.4 g L1 [6]. Finally, Inquil and Valverde achieved an efficiency of 79.58% in river water with a concentration of Pb 2.36 mg L1, 0.6 g L1 of nano-TiO2, and 1 h of interaction [41]. These results indicate that the equilibrium time estimated in this work is smaller than that reported by other authors (Table 4).
Therefore, it was shown that nano-TiO2 has high potential for Pb(II) adsorption by reducing the concentration from 47.54 mg L1 to 0.39 mg L1. Regarding the adsorbed amount, it was also constant from 1 h of interaction time, with the maximum adsorbed amount of 47.47 mg g1 at 15 h (see Figure 6).
The maximum adsorption capacity determined in this work was 65.99 mg g−1. This result is lower than the 194.28 mg g1 obtained by Kanna et al., where an initial concentration of Pb(II) of 400 mg L1, pH 7, and an adsorbent dose of 0.50 g L1, were used [42].
Table 4. Maximum capacities and adsorption percentages of different adsorbents towards Pb(II) in aqueous solutions.
Table 4. Maximum capacities and adsorption percentages of different adsorbents towards Pb(II) in aqueous solutions.
AdsorbentAdsorption Capacity
qm (mg g−1)
pHTime
(h)
Dose
(g L−1)
Ci
(mg L−1)
Reference
nano-TiO232.03820.010.1[43]
nano-TiO27.4164425[39]
nano-TiO2158.737240.32100[40]
Anatase194.2870.25 2 × 10 3 400[42]
Titania nanofiber2.56540.050.50[44]
Composed of pectin and nano-TiO266.285.51395.83[4]
Graphene oxide and TiO2 nanocomposite65.65.6120.02550[45]
Graphene oxide nanocomposite35.65.660.02550[45]
nano-ZnO6.76.67211010[46]
nano-CeO29.27NR2NR[47]
Modified nano-Al2O310051.5150[48]
Kaolinite7.754.548 0.12000[49]
Montmorillonite31.105.73 250[50]
Bentonite51.19NR320200[51]
Biomass of A. bisporus 33.785.043100[52]
Biomass of Aspergillus niger32.604.021100[53]
Anaerobic granular biomass255.004.0–5.50.510100[54]
nano-TiO265.996.51147.54This work.
NR: Indicates not reported.
In Table 4, a simple comparison of some adsorbents and nanoadsorbents for metal removal has been made. The adsorption capacity obtained in this work was higher than that of biomass-based adsorbents, except for anaerobic granular biomass [54]. Likewise, nanoadsorbents that have adsorption capacity values above nano-TiO2 are modified nano-Al2O3 and pectin-TiO2 nanocomposite. On the other hand, regarding the works carried out with nano-TiO2, Recillas et al. obtained an adsorption capacity of 158.73 mg g1, which may be higher due to the contact time and initial concentration conditions, which were 24 h and 100 mg L1, respectively [40]. Subsequently, Kanna et al. found an adsorption capacity of 194.28 mg g1 using anatase nanoparticles (neutral pH), a concentration of 400 mg L1, and a minimal dose of 2 × 103 g L−1. Remarkably, our work employed a similar minimum dose of adsorbent and achieved comparable results, demonstrating the effectiveness of this low-dose approach [42]. It is worth noting that the advantages of TiO2 NPs lie mainly in their high removal efficiency reported in previous studies, high specific surface area, low cost, availability, low environmental impact, and chemical activity, which means that modification of their adsorption capacity has been widely studied [5,19,55]. On the other hand, they should be considered as an eco-sustainable material because they can be reused by means of desorption. Evidence provided by Hu and Shipley indicates optimal desorption at pH 2 and up to eight desorption cycles; furthermore, an EDTA solution could be used for this purpose [56].
The kinetic behavior of Pb adsorption was investigated by analyzing the adsorption rate using linear and nonlinear PFO, PSO, E, and IDM, which are based on the adsorption equilibrium capacity (see Table S1 and Table 5). In this work, linear and nonlinear models were compared since it is known that linear models can generate errors as a consequence of transformations from nonlinear to linear forms, implying that they can be inaccurate. Therefore, we prefer to compare with the nonlinear forms to produce accurate parameter results [57].
Four different kinetic adsorption models were applied to examine the kinetic behavior of Pb(II) adsorption (Figure S4). These models are based on equilibrium adsorption capacity, which is used to analyze the adsorption rate. The PFO equation’s constant, k1, yielded a value of 0.001 h−1; however, the intercept was qe = 0.69 mg g−1, which was not the same as the value for qe found in the kinetic adsorption tests (qe = 47.15 mg g−1). Additionally, the R2 was close to 0, which further supports the lack of linearity. Therefore, it is concluded that the PFO, IDM, and E kinetic models cannot reproduce the results of kinetic adsorption. Instead, by applying the PSO model, the values k2 = 0.58 g mg−1 h−1 and qe = 46.29 mg g−1 were obtained, which correspond to the equilibrium adsorption capacity determined experimentally. The obtained R2 value of 0.99 affirms the nearly perfect linearity depicted in Figure S4b. This result suggests that the Pb(II) adsorption kinetics mechanism is governed by the PSO equation.
The PSO lineal fit model had a capacity of 46.29 mg g−1, equal to the value obtained by the nonlinear fit model. However, the velocity constant (k2) for the linear model was very low at 0.58 g mg−1 h−1 compared to the nonlinear fit of 12.74 g mg−1 h−1. Consequently, the linear model did not fit the experimental data well. On the other hand, the nonlinear models of PFO, PSO, and E presented a good fit to the experimental data with R2 greater than 0.92, except for the IDM model (see Figure 7). Therefore, in order to find the best fit, the BIC was calculated with Equation (S12) in the Supplementary Materials [58]. The PFO and PSO models yielded the lowest BIC values, with a difference between the two values of less than 2 (Table 5), so both models can represent the adsorption process of Pb on TiO2. In this regard, numerous authors have suggested that the PSO model best describes the adsorption procedure [39,59,60]; hence, the chemisorption process is the predominant one in the adsorption process.

2.3. Effect of pH

The pH is one of the most important adsorption parameters since it alters the charge on the adsorbent surface, causing a significant change in its adsorption capacity, in addition to affecting the existing form of Pb in the solution [61]. Figure 8 shows a low percentage of adsorption at low pH values of 2 to 3. In this range, the H+ ions are found with a higher concentration in the solution and mobility on the surface, generating competition with the Pb2+ ions for the adsorption sites, generating a decrease in the adsorption percentage [61]. The gradual increase in pH causes a decrease in H+ ions and an increase in hydroxyl groups on the nano-TiO2 surface due to the protonation-dissociation equilibrium [62]. This increases the active adsorption sites available for interaction with Pb2+ ions and promotes the electrostatic attraction and ion exchange that occurs between the positive charges of the metal ions and the negative charges on the adsorbent surface [63]; these lead ions can form species such as Pb(OH)+, Pb2(OH)3+, Pb3(OH)4+, and others [64].
Increasing the pH generated an increase in the adsorption efficiency, reaching a maximum percentage of adsorption at 6.5 with a value of 99.93%. In the same way, other studies obtained similar results for pH in the elimination of Pb with nano-TiO2 [27,39,40,56]. At pH values above the point of zero charge (pHPZC), for nano-TiO2 located at 6.5 [65], there was no significant reduction in the adsorption percentage since, with increasing pH, the nano-TiO2 surface increased its negative charge favoring adsorption [60]. However, the experiments were not conducted at pH levels greater than 8 due to the progressive precipitation of Pb and formation of complexes that prevented its adsorption [62].

2.4. Adsorbent Dose

Figure 9 illustrates the relationship between the Pb(II) removal efficiency and adsorbed amount against nano-TiO2 adsorbent dose. It was observed that the percentage of adsorption increased as the absorbent dose increased, given that, with a greater amount of adsorbent, the number of adsorption active sites on the nano-TiO2 surface increases and the competition between ions occupying the active sites reduces, as a result there is an increase in the adsorption of Pb ions [60,62]. On the other hand, the adsorbed amount decreases with increase in the dose since the amount of adsorbent is increased but the amount of solute in the solution remains constant [66]. This prevents reaching the maximum amount of adsorption in an active site. When the dose of the adsorbent coincides with the total amount of Pb2+ ions in the solution, the adsorbed amount reaches the highest possible value [59], in this case, 99.93% of Pb was removed from the water with a dose of 1 g L−1; Giammar et al. reported similar results [27]. After that, by further increasing the adsorption dose, the removal percentage decreased slightly.

2.5. Adsorption Isotherms

The adsorption isotherms obtained for the TiO2 NPs are plotted in Figure 10. An increase in adsorption capacity can be seen with increasing initial adsorbate concentration. The optimal conditions for obtaining the adsorption experimental data were a pH of 6.5, a dose of 1 g L−1, and 1 h of agitation at equilibrium. The adsorption parameters calculated from the fit, and the BIC and RSS values are presented in Table 6. Evidently, the Freundlich model yielded the lowest value of R2, so it was ruled out, implying that the process is not governed by multilayer. The other models had high values close to each other of R2. Therefore, another criterion by which to evaluate the models is the BIC value, where the lowest values were found for the Langmuir and Sips models; the difference between the BIC values of the two models was less than 2. In fact, both applied models could explain well the adsorption process of Pb(II) on nano-TiO2 [67]. The Langmuir model implies that the process is characterized by adsorption on the homogeneous surface sites of the adsorbent by the formation of a monolayer of adsorbate, with all sites being identical and having equivalent energies [39]. Likewise, the Sips model supports the Langmuir model—this was inferred from the value of ms = 0.85, close to unity, which confirms homogeneous adsorption [68]. It was also verified by the value of β = 0.97, close to 1, obtained by the Redlich–Peterson model, which indicates a monolayer reaction [69]. The maximum adsorption capacity predicted by the Langmuir model was 65.99 mg g−1, close to the value computed by the Sips model of 69.22 mg g−1. After comparing the adsorption capacity of TiO2, it was confirmed that it was superior to other adsorbents reported in the literature (Table 4).

2.6. Adsorption Mechanisms of Pb(II) on TiO2

High Pb adsorption was seen at values of pH above pHPZC = 6.5, because the adsorbent surface was negatively charged, which generated an increase in the Coulomb interaction between the ions and the adsorbent surface. In their study, Chen et al. showed how H2O adsorbs to TiO2 by bond breaking and formation, which is referred to as chemisorption. Their system demonstrated how the (O-H) bond in water dissociated and created new bonds that connected the H/O atom of H2O and the O/Ti atom of TiO2, which included forming the O-H bond on the surface of TiO2 [70]. Therefore, the ion exchange mechanism described by Equations (2) and (3) can be used to explain the adsorption of Pb(II) on nano-TiO2 as the binding of Pb2+, Pb(OH)+, and Pb(OH)2 with the active sites available on the surface of the adsorbents.
2 R O H + P b 2 + 2 R O P b + 2 H + Step 1
R O H + P b O H + R O P b O H + H + Step 2
or by hydrogen bonding given by Equation (4)
2 R O H + P b ( O H ) 2 R O H 2 + P b ( O H ) 2 Step 3
Since the oxygen of the (O-H) on the adsorbent’s surface is a strong Lewis base by its characteristic vacant doublet electrons, the oxygen forms a coordination complex with the low electron chemical species, such as metal ions [60,71], as shown in Figure 11. Step 1 and step 2 reveal the ion exchange first stage (deprotonation) and the adsorption of Pb2+ and Pb(OH)+ on the deprotonated surface active sites, respectively, expressed in Equations (2) and (3). Step 3 is the formation of hydrogen bonds, expressed in Equation (4).

3. Perspectives

For further studies, we intend to continue with the application of nano-TiO2 in water bodies with natural organic matter (NOM) since it has been found that nano-TiO2 with ions such as Pb2+ and Ca2+ adsorbed provides more complexation sites, which may favor the adsorption of NOM through the formation of bridges between the nano-TiO2 and NOM surface [72,73].
On the other hand, in the search to optimize the equilibrium time and the acceleration of the adsorption kinetics, it is intended to modify the surface properties of nano-TiO2 by adding materials such as graphene oxide [45], pectin [4], chitosan [74], carbon nanotubes, montmorillonite, binary oxides, and ternary compounds with magnetite and silica [75], whose adsorbents have a higher adsorption capacity, reaching equilibrium times of less than 1 h.
Considering that, after the adsorption process, the nano-TiO2 with Pb(II) ion species on its surface cannot be released into the environment due to its toxicity, the reusability of the nano-TiO2 and the recovery of the adsorbed metal are important, making the adsorption process more cost-effective and environmentally friendly. Hu and Shipley [56] investigated the regeneration capacity of nano-TiO2 for Pb(II) recovery using various salts, 95% of desorbed Pb(II) being obtained by use of NaNO3. On the other hand, EDTA was used for optimal regeneration of nano-TiO2, which can be reused even for removal in systems with various metals [56,76]. According to Engates and Shipley [43], nano-TiO2 could be reused in eight cycles of Pb(II) adsorption, in which a constant and similar adsorption capacity was achieved in each cycle. With respect to the use of desorbed Pb(II) as Pb(OH)2, this has applications in the field of chemical catalysis [77].

4. Materials and Methods

4.1. Materials and Chemicals

All the chemical reagents utilized were of analytical quality. TiO2 NPs were acquired from Sigma Aldrich (Darmstadt, Germany) and used unpurified. SCP Science (Quebec, QC, Canada) provided a Pb(II) standard solution with a concentration of 999 ± 3 µg mL−1. Merck (Darmstadt, Germany) brand hydrochloric acid (HCl) and sodium hydroxide (NaOH) reagents were used.
Water polluted with Pb was collected from the irrigation canal on the left bank of the Mantaro river (CIMIRM) at El Mantaro, Jauja, Peru, at 3320 m above sea level and 11°41′21.5″S 75°22′24.5″W. The original Pb(II) concentration of 0.0178 mg L−1 in the water was modified to 47.54 mg L−1. In addition, other heavy metals were found in the water, whose concentrations were compared to the maximum concentration established (presented in parentheses) by the Food and Agriculture Organization (FAO) for irrigation water [78]: As: 0.0176 mg L−1 (0.1 mg L−1), Cu: 0.021 mg L−1 (0.2 mg L−1), Ni: 0.0032 mg L−1 (0.2 mg L−1), Cd: 0.00047 mg L−1 (0.01 mg L−1), Cr: 0.0026 mg L−1 (0.1 mg L−1), Zn: 0.1305 mg L−1 (2 mg L−1) and Fe: 2.8477 mg L−1 (5 mg L−1). As can be seen, there were no concentrations above the limits established by FAO.

4.2. Characterization of Nano-TiO2

4.2.1. X-ray Diffraction (XRD)

XRD was used to characterize the materials using an Empyrean diffractometer with CuKα radiation at wavelength = 1.5406 Å. The X-ray diffractograms were taken following a step scanning arrangement with a range of 2θ = 20–80° and a resolution of 0.01° per step. Match V3 software was used to identify the crystallographic phase of anatase nano-TiO2, with the tetragonal crystalline structure, space group I41/amd, and cell parameters a = 3.7892 Å and c = 9.5370 Å as an initial matching candidate, with the crystallographic information files (CIF) #5000223 for the anatase phase. The peak’s profile function Thompson–Cox–Hastings pseudo-Voigt axial divergence asymmetry was used for the Rietveld refinement using the program FullProf Suite (version January 2021). Furthermore, the aluminum oxide Al2O3 standard was used to determine the diffractometer’s instrumental resolution function (IRF), which yielded Caglioti coefficients including U = 0.0093, V = 0.0051, and W = 0.0013 [31].

4.2.2. Scanning Electron Microscopy (SEM), Energy Dispersive Spectroscopy (EDS), and Transmission Electron Microscopy (TEM)

SEM and TEM techniques were employed to estimate the mean size of particles and the TiO2 morphologies. To assess the atomic composition, the elemental compositions of the materials were examined using EDS mapping.

4.2.3. X-ray Photoelectron Spectroscopy (XPS)

Chemical surface examinations on the studied materials were carried out utilizing the SPECS PHOIBOS 100/150 equipment with a hemispheric analyzer spectrometer, operating at 1486.6 eV of Al Kα. The XPS spectra were obtained using a high-resolution polychromatic X-ray source with a 0.02 eV energy step. The SPECS company’s Casa-XPS software was used to change the peak locations of the Ti 2p, O 1s, and Pb 4f levels to compute the chemical binding energies (BE) of produced species, as well as the relative atomic amounts on the sample surfaces. The spectra were calibrated using adventitious carbon (B.E. reference) at C 1 s = 284.6 eV after being calibrated with an electron flood cannon at 12 µA and 1 eV.

4.3. Pb(II) Removal Experiments with Nano-TiO2

The Pb(II) removal experiments were carried out in real water samples contaminated with Pb, taken from the CIMIRM-Junín-Perú irrigation canal, whose initial concentration was modified from 0.0178 mg L−1 to 47.54 mg L−1. To modify the concentration of this metal, a dilution was carried out from a standard solution of Pb at an initial concentration of 999 ± 3 µg mL−1. Pb quantification, in all cases, was performed using the inductively coupled plasma mass spectrometry (ICP-MS) analysis technique.

4.3.1. Adsorption Kinetics

Pb(II) adsorption kinetics experiments with nano-TiO2 were performed in order to determine the equilibrium time and evaluate the kinetic models.
Ten experiments were carried out depending on the contact time, in a range from 1 to 27 h, at 47.54 mg L−1 of initial concentration of Pb(II), using a nano-TiO2 dose of 1 g L−1, at pH 5.5, under constant stirring at 300 rpm, at room temperature.

4.3.2. pH Effect

In order to know the effect of pH on the removal of Pb(II) using nano-TiO2, experiments were carried out controlling pH in the range of 2 to 8, for a stirring time of 1 h (equilibrium time determined by the experiments of adsorption kinetics), with an adsorbent dose of 1 g L−1, maintaining constant stirring at 300 rpm at room temperature. The control of the previously indicated pH range was performed by adding drops of HCl and NaOH solutions in concentrations of 0.05 M, 0.1 M, 1 M, and 4 M.

4.3.3. Adsorbent Dose

The adsorbent dose experiments were carried out on real water samples at an initial Pb(II) concentration of 47.54 mg L−1. The nano-TiO2 dose range considered in this work was from 0.2 g L−1 to 1.5 g L−1. The other experimental parameters were: pH 6.5, constant stirring time of 1 h at 300 rpm at room temperature.

4.3.4. Adsorption Isotherms

Pb(II) adsorption isotherms with nano-TiO2 were constructed using the previously determined experimental conditions: the equilibrium time, the optimum pH, and the adsorbent dose. Utilizing these values, ten tests were conducted on irrigation canal water samples with beginning concentrations ranging from 3 to 65 mg L−1. To analyze the data obtained, the Langmuir, Freundlich, Redlich–Peterson, Sips and Temkin adsorption models were used.
The theoretical description of the kinetic and isotherm models is given in the Supplementary Materials.

5. Conclusions

The removal efficiency of Pb(II) using nano-TiO2 in irrigation water was evaluated at a modified initial concentration of 47.54 mg L−1, in the presence of other metals. This nanomaterial was characterized by XRD after adsorption, confirming the presence of the tetragonal crystalline structure of the anatase phase. The mean crystallite size was obtained by Rietveld refinement giving a value of 9.9 nm; the mean particle size was determined by TEM, being 22.46 nm. The XPS studies indicated that there was no alteration of the absorption bands with respect to increase in the concentration of Pb(II). The SEM images showed changes in the surface morphology of the nano-TiO2, confirming the occurrence of metal adsorption. In the Pb(II) removal experiments, more than 99% Pb(II) elimination was achieved by determining the optimal contact time, pH, and dose of adsorbent, having values of 1 h, 6.5, and 1 g L−1, suggesting that there was no significant influence of the prevalence of other metal ions. The experimental data were evaluated with linear and nonlinear models of adsorption kinetics. The best fit was found to correspond to the PSO model, indicating that the process was governed by chemisorption. This result was supported by evaluation of the adsorption isotherms, where the Langmuir and Sips models provided the best fit, evidencing that adsorption occurred in the homogeneous sites of the nano-TiO2 surface through the formation of a adsorbate monolayer. In addition, a high adsorption capacity of nano-TiO2 for Pb(II) of 65.99 mg g−1 was estimated, higher than the values reported in other studies. In general, the results of this study indicate that nano-TiO2 is a potential adsorbent for application in the treatment of water contaminated with Pb(II).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124596/s1, Supporting section: Theoretical background of kinetic and isotherm adsorption models. Figure S1: (a,f) SEM images, (b,g) EDS mapping images, and (c,d,e,h,i,j) elemental EDS images for nano-TiO2 with C0 (Pb(II)) = 1.16 mg L−1. Figure S2: EDS spectrum for both nano-TiO2 samples exposed to two different Pb(II) concentrations. Figure S3: (a) TEM image of TiO2-PbC2 NPs (bar length of 100 nm), (b) PSD for nano-TiO2-PbC2, (c) Zoomed TEM image of nano-TiO2-PbC2, (d) SAED pattern of nano-TiO2-PbC2, and (e) crystallite size distribution histogram for nano-TiO2-PbC2 obtained from Rietveld refinement. C0 (Pb(II)) = 1.16 mg L−1. Figure S4: Linear models of Pb (II) adsorption kinetics on nano-TiO2. Adsorbent dose of 1 g L−1, pH 5.5. Co = 47.54 mg L−1. Table S1: Fit parameters of linear models of adsorption kinetics of PFO, PSO, E, and IDM. References [79,80,81,82,83,84,85,86,87,88,89,90] are cited in Supplementary Materials.

Author Contributions

Conceptualization, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; methodology, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; software, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; validation, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; formal analysis, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; investigation, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; resources, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; data curation, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; writing—original draft preparation, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; writing—review and editing, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; visualization, M.A.V.-C., Y.C.-H., A.F.M.-A., J.P.-V., N.-R.C.-H., Y.B.-R. and J.A.R.-G.; supervision, Y.B.-R. and J.A.R.-G.; project administration, Y.B.-R. and J.A.R.-G.; funding acquisition, J.A.R.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Vicerrectorado de Investigación y Posgrado (VRIP) de la Universidad Nacional Mayor de San Marcos (UNMSM)—RR N° 09412-R-21 and project number B2113002i-MINEDU-2021. The APC was funded by VRIP-UNMSM.

Data Availability Statement

The original data related to this research can be obtained at any time via the corresponding author’s email: [email protected].

Acknowledgments

We thank the Investigación y Posgrado (VRIP) de la Universidad Nacional Mayor de San Marcos (UNMSM)—RR N° 09412-R-21 and project number B2113002i-MINEDU-2021 for financially supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Lu, F.; Astruc, D. Nanomaterials for removal of toxic elements from water. Coord. Chem. Rev. 2018, 356, 147–164. [Google Scholar] [CrossRef]
  2. Baruah, S.; Najam Khan, M.; Dutta, J. Perspectives and applications of nanotechnology in water treatment. Environ. Chem. Lett. 2016, 14, 1–14. [Google Scholar] [CrossRef]
  3. World Water Development Report 2018|UN-Water. Available online: https://www.unwater.org/publications/world-water-development-report-2018 (accessed on 19 March 2023).
  4. Bok-Badura, J.; Jakóbik-Kolon, A.; Karoń, K.; Mitko, K. Sorption studies of heavy metal ions on pectin-nano-titanium dioxide composite adsorbent. Sep. Sci. Technol. 2018, 53, 1034–1044. [Google Scholar] [CrossRef]
  5. Bhatia, M.; Satish Babu, R.; Sonawane, S.H.; Gogate, P.R.; Girdhar, A.; Reddy, E.R.; Pola, M. Application of nanoadsorbents for removal of lead from water. Int. J. Environ. Sci. Technol. 2017, 14, 1135–1154. [Google Scholar] [CrossRef]
  6. Jyothi, B.; Singa, P.M.; Desai, N.N. Performance of TiO2 Nano powder in Removal of Lead from Synthetic Solution. Int. Res. J. Eng. Technol. 2015, 2, 1396–1399. [Google Scholar]
  7. Sepehri, S.; Kanani, E.; Abdoli, S.; Rajput, V.D.; Minkina, T.; Asgari Lajayer, B. Pb(II) Removal from Aqueous Solutions by Adsorption on Stabilized Zero-Valent Iron Nanoparticles—A Green Approach. Water 2023, 15, 222. [Google Scholar] [CrossRef]
  8. Chanthapon, N.; Sarkar, S.; Kidkhunthod, P.; Padungthon, S. Lead removal by a reusable gel cation exchange resin containing nano-scale zero valent iron. Chem. Eng. J. 2018, 331, 545–555. [Google Scholar] [CrossRef]
  9. Wang, C.; Wang, H.; Cao, Y. Pb(II) sorption by biochar derived from Cinnamomum camphora and its improvement with ultrasound-assisted alkali activation. Colloids Surf. A Physicochem. Eng. Asp. 2018, 556, 177–184. [Google Scholar] [CrossRef]
  10. Moustafa, H.; Isawi, H.; Abd El Wahab, S.M. Utilization of PVA nano-membrane based synthesized magnetic GO-Ni-Fe2O4 nanoparticles for removal of heavy metals from water resources. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100696. [Google Scholar] [CrossRef]
  11. World Health Organization. Guidelines for Drinking-Water Quality; World Health Organization: Geneva, Switzerland, 2011; ISBN 9789241548151. [Google Scholar]
  12. Long, Y.; Jiang, J.; Hu, J.; Hu, X.; Yang, Q.; Zhou, S. Removal of Pb(II) from aqueous solution by hydroxyapatite/carbon composite: Preparation and adsorption behavior. Colloids Surf. A Physicochem. Eng. Asp. 2019, 577, 471–479. [Google Scholar] [CrossRef]
  13. Van Tran, C.; Quang, D.V.; Nguyen Thi, H.P.; Truong, T.N.; La, D.D. Effective Removal of Pb(II) from Aqueous Media by a New Design of Cu-Mg Binary Ferrite. ACS Omega 2020, 5, 7298–7306. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, Y.Y.; Liu, Y.X.; Lu, H.H.; Yang, R.Q.; Yang, S.M. Competitive adsorption of Pb(II), Cu(II), and Zn(II) ions onto hydroxyapatite-biochar nanocomposite in aqueous solutions. J. Solid State Chem. 2018, 261, 53–61. [Google Scholar] [CrossRef]
  15. Carolin, C.F.; Kumar, P.S.; Saravanan, A.; Joshiba, G.J.; Naushad, M. Efficient techniques for the removal of toxic heavy metals from aquatic environment: A review. J. Environ. Chem. Eng. 2017, 5, 2782–2799. [Google Scholar] [CrossRef]
  16. Azimi, A.; Azari, A.; Rezakazemi, M.; Ansarpour, M. Removal of Heavy Metals from Industrial Wastewaters: A Review. Chem. Bio. Eng. Rev. 2017, 4, 37–59. [Google Scholar] [CrossRef]
  17. Bolisetty, S.; Peydayesh, M.; Mezzenga, R. Sustainable technologies for water purification from heavy metals: Review and analysis. Chem. Soc. Rev. 2019, 48, 463–487. [Google Scholar] [CrossRef] [PubMed]
  18. Lata, S.; Singh, P.K.; Samadder, S.R. Regeneration of adsorbents and recovery of heavy metals: A review. Int. J. Environ. Sci. Technol. 2015, 12, 1461–1478. [Google Scholar] [CrossRef] [Green Version]
  19. Santhosh, C.; Velmurugan, V.; Jacob, G.; Jeong, S.K.; Grace, A.N.; Bhatnagar, A. Role of nanomaterials in water treatment applications: A review. Chem. Eng. J. 2016, 306, 1116–1137. [Google Scholar] [CrossRef]
  20. Hashem, M.; Al-Karagoly, H. Synthesis, characterization, and cytotoxicity of titanium dioxide nanoparticles and in vitro study of its impact on lead concentrations in bovine blood and milk. J. Biotech Res. 2021, 12, 93–105. [Google Scholar]
  21. Motlochová, M.; Slovák, V.; Pližingrová, E.; Lidin, S.; Šubrt, J. Highly-efficient removal of Pb(II), Cu(II) and Cd(II) from water by novel lithium, sodium and potassium titanate reusable microrods. RSC Adv. 2020, 10, 3694–3704. [Google Scholar] [CrossRef] [Green Version]
  22. Govan, J. Recent advances in magnetic nanoparticles and nanocomposites for the remediation of water resources. Magnetochemistry 2020, 6, 49. [Google Scholar] [CrossRef]
  23. Hornyak, G.L. Introduction to Nanoscience and Nanotechnology; CRC Press: Boca Raton, FL, USA, 2008; ISBN 9781420048056. [Google Scholar]
  24. Zhu, Y.; Liu, X.; Hu, Y.; Wang, R.; Chen, M.; Wu, J.; Wang, Y.; Kang, S.; Sun, Y.; Zhu, M. Behavior, remediation effect and toxicity of nanomaterials in water environments. Environ. Res. 2019, 174, 54–60. [Google Scholar] [CrossRef]
  25. George, R.; Bahadur, N.; Singh, N.; Singh, R.; Verma, A.; Shukla, A.K. Environmentally Benign TiO2 Nanomaterials for Removal of Heavy Metal Ions with Interfering Ions Present in Tap Water. Mater. Today Proc. 2016, 3, 162–166. [Google Scholar] [CrossRef]
  26. Song, H.; Kuang, X.; Wei, X.; Luo, S.; Zeng, Q.; Peng, L. The effect of TiO2 nanoparticles size on Cd (II) removal by the paddy crusts from waterbody. J. Environ. Chem. Eng. 2022, 10, 107883. [Google Scholar] [CrossRef]
  27. Giammar, D.E.; Maus, C.J.; Xie, L. Effects of particle size and crystalline phase on lead adsorption to titanium dioxide nanoparticles. Environ. Eng. Sci. 2007, 24, 85–95. [Google Scholar] [CrossRef]
  28. Prasai, B.; Cai, B.; Underwood, M.K.; Lewis, J.P.; Drabold, D.A. Properties of amorphous and crystalline titanium dioxide from first principles. J. Mater. Sci. 2012, 47, 7515–7521. [Google Scholar] [CrossRef]
  29. Custodio, M.; Álvarez, D.; Cuadrado, W.; Montalvo, R.; Ochoa, S. Potentially toxic metals and metalloids in surface water intended for human consumption and other uses in the Mantaro River watershed, Peru. Soil Water Res. 2020, 15, 237–245. [Google Scholar] [CrossRef] [Green Version]
  30. Popa, N.C. The (hkl) dependence of diffraction-line broadening caused by strain and size for all laue groups in rietveld refinement. J. Appl. Crystallogr. 1998, 31, 176–180. [Google Scholar] [CrossRef]
  31. Canchanya-Huaman, Y.; Mayta-Armas, A.F.; Pomalaya-Velasco, J.; Bendezú-Roca, Y.; Guerra, J.A.; Ramos-Guivar, J.A. Strain and grain size determination of CeO2 and TiO2 nanoparticles: Comparing integral breadth methods versus Rietveld, µ-Raman, and TEM. Nanomaterials 2021, 11, 2311. [Google Scholar] [CrossRef]
  32. Patidar, V.; Jain, P. Green Synthesis of TiO2 Nanoparticle Using Moringa Oleifera Leaf Extract. Int. Res. J. Eng. Technol. 2017, 4, 470–473. [Google Scholar]
  33. Thamaphat, K.; Limsuwan, P.; Ngotawornchai, B. Phase Characterization of TiO2 Powder by XRD and TEM. Nat. Sci. 2008, 42, 357–361. [Google Scholar]
  34. Zhang, S.; Shi, Q.; Chou, T.M.; Christodoulatos, C.; Korfiatis, G.P.; Meng, X. Mechanistic study of Pb(II) removal by TiO2 and effect of PO4. Langmuir 2020, 36, 13918–13927. [Google Scholar] [CrossRef]
  35. Dementjev, A.P.; Ivanova, O.P.; Vasilyev, L.A.; Naumkin, A.V.; Nemirovsky, D.M.; Shalaev, D.Y. Altered layer as sensitive initial chemical state indicator. J. Vac. Sci. Technol. A Vac. Surf. Film. 1994, 12, 423–427. [Google Scholar] [CrossRef]
  36. Lu, C.J.; Kuang, A.X.; Huang, G.Y. X-ray photoelectron spectroscopy study on composition and structure of sol-gel derived PbTiO3 thin films. J. Appl. Phys. 1996, 80, 202–206. [Google Scholar] [CrossRef]
  37. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of XPS Data; Perkin-Elmer Corporation: Waltham, MA, USA, 1992. [Google Scholar]
  38. Beamson, G.; Briggs, D. High Resolution XPS of Organic Polymers: The Scienta ESCA300 Database. J. Chem. Educ. 1993, 70, 778. [Google Scholar] [CrossRef]
  39. Poursani, A.S.; Nilchi, A.; Hassani, A. The synthesis of nano TiO2 and its use for removal of lead ions from aqueous solution. J. Water Resour. Prot. 2016, 08, 438–448. [Google Scholar] [CrossRef] [Green Version]
  40. Recillas, S.; García, A.; González, E.; Casals, E.; Puntes, V.; Sánchez, A.; Font, X. Use of CeO2, TiO2 and Fe3O4 nanoparticles for the removal of lead from water. Toxicity of nanoparticles and derived compounds. Desalination 2011, 277, 213–220. [Google Scholar] [CrossRef] [Green Version]
  41. Inquil Ayquipa, L.; Valverde Flores, J. Adsorption of lead and iron present in the waters of the Santa River using titanium dioxide nanoparticles (TiO2). J. Nanotechnol. 2020, 4, 8–10. [Google Scholar] [CrossRef]
  42. Kanna, M.; Wongnawa, S.; Sherdshoopongse, P. Adsorption behavior of some metal ions on hydrated amorphous titanium dioxide surface. Songklanakarin J. Sci. Technol. 2005, 27, 1017–1026. [Google Scholar]
  43. Engates, K.E.; Shipley, H.J. Adsorption of Pb, Cd, Cu, Zn, and Ni to titanium dioxide nanoparticles: Effect of particle size, solid concentration, and exhaustion. Environ. Sci. Pollut. Res. 2011, 18, 386–395. [Google Scholar] [CrossRef]
  44. Dixit, A.; Mishra, P.K.; Alam, M.S. Titania Nanofibers: A Potential Adsorbent for Mercury and Lead Uptake. Int. J. Chem. Eng. Appl. 2017, 8, 75–81. [Google Scholar] [CrossRef] [Green Version]
  45. Lee, Y.C.; Yang, J.W. Self-assembled flower-like TiO2 on exfoliated graphite oxide for heavy metal removal. J. Ind. Eng. Chem. 2012, 18, 1178–1185. [Google Scholar] [CrossRef]
  46. Kikuchi, Y.; Qian, Q.; Machida, M.; Tatsumoto, H. Effect of ZnO loading to activated carbon on Pb(II) adsorption from aqueous solution. Carbon 2006, 44, 195–202. [Google Scholar] [CrossRef]
  47. Cao, C.Y.; Cui, Z.M.; Chen, C.Q.; Song, W.G.; Cai, W. Ceria hollow nanospheres produced by a template-free microwave-assisted hydrothermal method for heavy metal ion removal and catalysis. J. Phys. Chem. C 2010, 114, 9865–9870. [Google Scholar] [CrossRef]
  48. Afkhami, A.; Saber-Tehrani, M.; Bagheri, H. Simultaneous removal of heavy-metal ions in wastewater samples using nano-alumina modified with 2,4-dinitrophenylhydrazine. J. Hazard Mater. 2010, 181, 836–844. [Google Scholar] [CrossRef]
  49. Shahmohammadi-Kalalagh, S.; Babazadeh, H.; Manshouri, M. Isotherm and kinetic studies on adsorption of Pb, Zn and Cu by Kaolinite. Casp. J. Environ. Sci. 2010, 9, 243–255. [Google Scholar]
  50. Sen Gupta, S.; Bhattacharyya, K.G. Immobilization of Pb(II), Cd(II) and Ni(II) ions on kaolinite and montmorillonite surfaces from aqueous medium. J. Environ. Manag. 2008, 87, 46–58. [Google Scholar] [CrossRef] [PubMed]
  51. Al-Jlil, S.A. Kinetic study of adsorption of chromium and lead ions on bentonite clay using novel internal series model. Trends Appl. Sci. Res. 2015, 10, 38–53. [Google Scholar] [CrossRef] [Green Version]
  52. Vimala, R.; Das, N. Biosorption of cadmium (II) and lead (II) from aqueous solutions using mushrooms: A comparative study. J. Hazard Mater. 2009, 168, 376–382. [Google Scholar] [CrossRef]
  53. Dursun, A.Y. A comparative study on determination of the equilibrium, kinetic and thermodynamic parameters of biosorption of copper(II) and lead(II) ions onto pretreated Aspergillus niger. Biochem. Eng. J. 2006, 28, 187–195. [Google Scholar] [CrossRef]
  54. Hawari, A.H.; Mulligan, C.N. Biosorption of lead(II), cadmium(II), copper(II) and nickel(II) by anaerobic granular biomass. Bioresour. Technol. 2006, 97, 692–700. [Google Scholar] [CrossRef]
  55. Singh, S.; Barick, K.C.; Bahadur, D. Functional oxide nanomaterials and nanocomposites for the removal of heavy metals and dyes. Nanomater. Nanotechnol. 2013, 3, 3–20. [Google Scholar] [CrossRef] [Green Version]
  56. Hu, J.; Shipley, H.J. Regeneration of spent TiO2 nanoparticles for Pb (II), Cu(II), and Zn(II) removal. Environ. Sci. Pollut. Res. 2013, 20, 5125–5137. [Google Scholar] [CrossRef]
  57. López-Luna, J.; Ramírez-Montes, L.E.; Martinez-Vargas, S.; Martínez, A.I.; Mijangos-Ricardez, O.F.; González-Chávez, M.d.C.A.; Carrillo-González, R.; Solís-Domínguez, F.A.; Cuevas-Díaz, M.d.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. [Google Scholar] [CrossRef] [Green Version]
  58. Cao, Y.; Dong, S.; Dai, Z.; Zhu, L.; Xiao, T.; Zhang, X.; Yin, S.; Soltanian, M.R. Adsorption model identification for chromium(VI) transport in unconsolidated sediments. J. Hydrol. 2021, 598, 126228. [Google Scholar] [CrossRef]
  59. Zhao, X.; Jia, Q.; Song, N.; Zhou, W.; Li, Y. Adsorption of Pb(II) from an aqueous solution by titanium dioxide/carbon nanotube nanocomposites: Kinetics, thermodynamics, and isotherms. J. Chem. Eng. Data 2010, 55, 4428–4433. [Google Scholar] [CrossRef]
  60. Mostafa, N.G.; Yunnus, A.F.; Elawwad, A. Adsorption of Pb(II) from water onto ZnO, TiO2, and Al2O3: Process study, adsorption behaviour, and thermodynamics. Adsorpt. Sci. Technol. 2022, 2022, 7582756. [Google Scholar] [CrossRef]
  61. Shu, Y.; Huang, R.; Wei, X.; Liu, L.; Jia, Z. Pb(II) Removal using TiO2-Embedded Monolith Composite Cryogel as an alternative wastewater treatment method. Water Air Soil Pollut. 2017, 228, 375. [Google Scholar] [CrossRef]
  62. Karapinar, H.S.; Kilicel, F.; Ozel, F.; Sarilmaz, A. Fast and effective removal of Pb(II), Cu(II) and Ni(II) ions from aqueous solutions with TiO2 nanofibers: Synthesis, adsorption-desorption process and kinetic studies. Int. J. Environ. Anal. Chem. 2021, 103, 1–22. [Google Scholar] [CrossRef]
  63. Maleki, A.; Hayati, B.; Najafi, F.; Gharibi, F.; Joo, S.W. Heavy metal adsorption from industrial wastewater by PAMAM/TiO2 nanohybrid: Preparation, characterization and adsorption studies. J. Mol. Liq. 2016, 224, 95–104. [Google Scholar] [CrossRef]
  64. Tao, Y.; Ye, L.; Pan, J.; Wang, Y.; Tang, B. Removal of Pb(II) from aqueous solution on chitosan/TiO2 hybrid film. J. Hazard Mater. 2009, 161, 718–722. [Google Scholar] [CrossRef]
  65. Mendoza-Villa, F.; Checca-Huaman, N.R.; Ramos-Guivar, J.A. Ecotoxicological properties of titanium dioxide nanomorphologies in daphnia magna. Nanomaterials 2023, 13, 927. [Google Scholar] [CrossRef] [PubMed]
  66. Bhargavi, R.J.; Maheshwari, U.; Gupta, S. Synthesis and use of alumina nanoparticles as an adsorbent for the removal of Zn(II) and CBG dye from wastewater. Int. J. Ind. Chem. 2014, 6, 31–41. [Google Scholar] [CrossRef] [Green Version]
  67. Lima, E.C.; Sher, F.; Guleria, A.; Reza, M.; Anastopoulos, I.; Nguyen, H.; Hosseini-bandegharaei, A. Is One Performing the Treatment Data of Adsorption Kinetics Correctly? J. Environ. Chem. Eng. 2021, 9, 104813. [Google Scholar] [CrossRef]
  68. Ramos Guivar, J.A.; Bustamante, D.A.; Gonzalez, J.C.; Sanches, E.A.; Morales, M.A.; Raez, J.M.; López-Muñoz, M.J.; Arencibia, A. Adsorption of arsenite and arsenate on binary and ternary magnetic nanocomposites with high iron oxide content. Appl. Surf. Sci. 2018, 454, 87–100. [Google Scholar] [CrossRef]
  69. Razzaz, A.; Ghorban, S.; Hosayni, L.; Irani, M.; Aliabadi, M. Chitosan nanofibers functionalized by TiO2 nanoparticles for the removal of heavy metal ions. J. Taiwan Inst. Chem. Eng. 2016, 58, 333–343. [Google Scholar] [CrossRef]
  70. Chen, B.; Li, L.; Liu, L.; Cao, J. Molecular simulation of adsorption properties of thiol-functionalized titanium dioxide (TiO2) nanostructure for heavy metal ions removal from aqueous solution. J. Mol. Liq. 2022, 346, 118281. [Google Scholar] [CrossRef]
  71. Taty-Costodes, V.C.; Fauduet, H.; Porte, C.; Delacroix, A. Removal of Cd(II) and Pb(II) ions, from aqueous solutions, by adsorption onto sawdust of Pinus sylvestris. J. Hazard Mater. 2003, 105, 121–142. [Google Scholar] [CrossRef]
  72. Wang, D.; Wang, P.; Wang, C.; Ao, Y. Effects of interactions between humic acid and heavy metal ions on the aggregation of TiO2 nanoparticles in water environment. Environ. Pollut. 2019, 248, 834–844. [Google Scholar] [CrossRef]
  73. Ramirez, L.; Gentile, R.S.; Zimmerman, S.; Stoll, S. Behavior of TiO2 and CeO2 nanoparticles and polystyrene nanoplastics in bottled mineral, drinking and Lake Geneva Waters. Impact of water hardness and natural organic matter on nanoparticle surface properties and aggregation. Water 2019, 11, 721. [Google Scholar] [CrossRef] [Green Version]
  74. Spoială, A.; Ilie, C.-I.; Dolete, G.; Croitoru, A.-M.; Surdu, V.-A.; Trușcă, R.-D.; Motelica, L.; Oprea, O.-C.; Ficai, D.; Ficai, A.; et al. Preparation and characterization of chitosan/TiO2 composite membranes as adsorbent materials for water purification. Membranes 2022, 12, 804. [Google Scholar] [CrossRef]
  75. Ashraf, S.; Siddiqa, A.; Shahida, S.; Qaisar, S. Titanium-based nanocomposite materials for arsenic removal from water: A review. Heliyon 2019, 5, e01577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Hu, J.; Shipley, H.J. Evaluation of desorption of Pb(II), Cu(II) and Zn(II) from titanium dioxide nanoparticles. Sci. Total Environ. 2012, 431, 209–220. [Google Scholar] [CrossRef] [PubMed]
  77. Perera, W.N.; Hefter, G.; Sipos, P.M. An investigation of the lead (II)-Hydroxide System. Inorg. Chem. 2001, 40, 3974–3978. [Google Scholar] [CrossRef] [PubMed]
  78. Ayers, R.S.; Westcot, D.W. Water Quality for Agriculture; Food and Agriculture Organization of the United Nations: Rome, Italy, 1985; ISBN 92-5-102263-1. [Google Scholar]
  79. Ho, Y.S. Citation review of Lagergren kinetic rate equation on adsorption reactions. Scientometrics 2004, 59, 171–177. [Google Scholar] [CrossRef]
  80. Tseng, R.L.; Wu, F.C.; Juang, R.S. Characteristics and applications of the Lagergren’s first-order equation for adsorption kinetics. J. Taiwan Inst. Chem. Eng. 2010, 41, 661–669. [Google Scholar] [CrossRef]
  81. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  82. Hamidpour, M.; Hosseini, N.; Mozafari, V.; Heshmati, M. Removal of Cd(II) and Pb(II) from aqueous solutions by pistachio hull waste. Rev. Int. de Contam. 2018, 34, 307–316. [Google Scholar] [CrossRef] [Green Version]
  83. Wu, F.C.; Tseng, R.L.; Juang, R.S. Characteristics of Elovich equation used for the analysis of adsorption kinetics in dye-chitosan systems. Chem. Eng. J. 2009, 150, 366–373. [Google Scholar] [CrossRef]
  84. Wang, J.; Guo, X. Rethinking of the Intraparticle Diffusion Adsorption Kinetics Model: Interpretation, Solving Methods and Applications. Chemosphere 2022, 309, 1–14. [Google Scholar] [CrossRef] [PubMed]
  85. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef] [Green Version]
  86. Skopp, J. Derivation of the Freundlich Adsorption Isotherm from Kinetics. J. Chem. Educ. 2009, 86, 1341–1343. [Google Scholar] [CrossRef]
  87. Chu, K.H. Revisiting the Temkin Isotherm: Dimensional Inconsistency and Approximate Forms. Ind. Eng. Chem. Res. 2021, 60, 13140–13147. [Google Scholar] [CrossRef]
  88. Sips, R. On the structure of a catalyst surface. J. Chem. Phys. 1948, 16, 490–495. [Google Scholar] [CrossRef]
  89. Redlich, O.; Peterson, D.L. A useful adsorption isotherm. J. Phys. Chem. 1959, 63, 1024. [Google Scholar] [CrossRef]
  90. James, G.; Witten, D.; Hastie, T.; Tibshirani, R. An Introduction to Statistical Learning; Springer: New York, NY, USA, 2017. [Google Scholar] [CrossRef]
Figure 1. Rietveld refinement of X-ray diffractograms of TiO2 anatase after adsorption of Pb(II) for Co = 47.54 mg L−1 (a) and Pb Co = 1.16 mg L−1 (b). The black dots (Iobs) represent the experimental data, the red lines (Ical) are the calculated XRD diffractogram, the green vertical lines are the Bragg positions, and the blue lines represent the residual lines.
Figure 1. Rietveld refinement of X-ray diffractograms of TiO2 anatase after adsorption of Pb(II) for Co = 47.54 mg L−1 (a) and Pb Co = 1.16 mg L−1 (b). The black dots (Iobs) represent the experimental data, the red lines (Ical) are the calculated XRD diffractogram, the green vertical lines are the Bragg positions, and the blue lines represent the residual lines.
Molecules 28 04596 g001
Figure 2. (a,f) SEM images, (b,g) EDS mapping images, and (ce,hj) elemental EDS images for nano-TiO2 with C0 (Pb(II)) = 47.54 mg L−1.
Figure 2. (a,f) SEM images, (b,g) EDS mapping images, and (ce,hj) elemental EDS images for nano-TiO2 with C0 (Pb(II)) = 47.54 mg L−1.
Molecules 28 04596 g002
Figure 3. (a) TEM image of TiO2-PbC1 NPs (bar length of 100 nm), (b) PSD histogram for nano-TiO2-PbC1, (c) Zoomed TEM image of nano-TiO2-PbC1, (d) SAED pattern of nano-TiO2-PbC1, and (e) crystallite size distribution histogram for nano-TiO2-PbC1 obtained from Rietveld refinement. C0 (Pb(II)) = 47.54 mg L−1.
Figure 3. (a) TEM image of TiO2-PbC1 NPs (bar length of 100 nm), (b) PSD histogram for nano-TiO2-PbC1, (c) Zoomed TEM image of nano-TiO2-PbC1, (d) SAED pattern of nano-TiO2-PbC1, and (e) crystallite size distribution histogram for nano-TiO2-PbC1 obtained from Rietveld refinement. C0 (Pb(II)) = 47.54 mg L−1.
Molecules 28 04596 g003
Figure 4. XPS spectra of the TiO2-PbC1 sample, showing the deconvoluted peaks of: (a) Ti 2p, (b) Pb 4f, and (c) O 1s.
Figure 4. XPS spectra of the TiO2-PbC1 sample, showing the deconvoluted peaks of: (a) Ti 2p, (b) Pb 4f, and (c) O 1s.
Molecules 28 04596 g004
Figure 5. XPS spectra of the TiO2Pb-C2 sample, showing the deconvoluted peaks of: (a) Ti 2p, (b) Pb 4f, and (c) O 1s.
Figure 5. XPS spectra of the TiO2Pb-C2 sample, showing the deconvoluted peaks of: (a) Ti 2p, (b) Pb 4f, and (c) O 1s.
Molecules 28 04596 g005
Figure 6. Evaluation of time on the efficiency and adsorbed amount of Pb(II) onto nano-TiO2.
Figure 6. Evaluation of time on the efficiency and adsorbed amount of Pb(II) onto nano-TiO2.
Molecules 28 04596 g006
Figure 7. Nonlinear models of Pb(II) adsorption kinetics on nano-TiO2. Adsorbent dose of 1 g L−1, pH 5.5, Co = 47.54 mg L−1.
Figure 7. Nonlinear models of Pb(II) adsorption kinetics on nano-TiO2. Adsorbent dose of 1 g L−1, pH 5.5, Co = 47.54 mg L−1.
Molecules 28 04596 g007
Figure 8. pH dependence of the Pb removal on nano-TiO2.
Figure 8. pH dependence of the Pb removal on nano-TiO2.
Molecules 28 04596 g008
Figure 9. Adsorbent dose dependence of the Pb removal and adsorbed amount for nano-TiO2, equilibrium time of 1 h.
Figure 9. Adsorbent dose dependence of the Pb removal and adsorbed amount for nano-TiO2, equilibrium time of 1 h.
Molecules 28 04596 g009
Figure 10. Pb(II) adsorption isotherms from irrigation water using nano-TiO2.
Figure 10. Pb(II) adsorption isotherms from irrigation water using nano-TiO2.
Molecules 28 04596 g010
Figure 11. Pb(II) adsorption mechanisms on nano-TiO2.
Figure 11. Pb(II) adsorption mechanisms on nano-TiO2.
Molecules 28 04596 g011
Table 1. Rietveld refinement parameters of the TiO2-PbC1 and TiO2-PbC2 samples: lattice parameters, cell volume, and Caglioti values. Agreement R-factors Rexp (%), Rp (%), Rwp (%), and goodness of fit, chi-square (χ2).
Table 1. Rietveld refinement parameters of the TiO2-PbC1 and TiO2-PbC2 samples: lattice parameters, cell volume, and Caglioti values. Agreement R-factors Rexp (%), Rp (%), Rwp (%), and goodness of fit, chi-square (χ2).
Refined ParametersTiO2-PbC1
Co = 47.54 mg L−1
TiO2-PbC2
Co = 1.16 mg L−1
a ( ) 3.782 (9)3.798 (9)
b ( ) 3.782 (9)3.798 (9)
c ( ) 9.488 (3)9.527 (3)
α ( ° ) 9090
β ( ° ) 9090
γ ( ° ) 9090
V ( 3 ) 135.750 (6)137.441 (6)
Y 00 −0.077 (7)0.022 (4)
Y 20 0.675 (8)0.972 (8)
Y 40 0.650 (9)0.180 (9)
Y 44 + 0.045 (6)−0.127 (5)
Y 44 −0.958 (2)−0.585 (2)
Y 60 −0.120 (9)0.554 (9)
Y 64 + −0.368 (4)−0.184 (3)
Y 64 −2.311 (2)0.952 (2)
FWHM parameters
U4.4644.342
V−2.569−2.549
W0.9070.919
Average max strain108.48 (4)98.63 (3)
Average size (nm)9.9 (7)8.7 (5)
R e x p ( % ) 4.524.54
R p ( % ) 7.216.90
R w p ( % ) 7.707.38
χ22.902.63
Table 2. Elemental composition for the studied samples obtained from EDS spectrum analysis.
Table 2. Elemental composition for the studied samples obtained from EDS spectrum analysis.
SampleTi (% wt)O (% wt)Pb (% wt)
TiO2-PbC147.7 ± 0.247.5 ± 0.24.8 ± 0.1
TiO2-PbC257.8 ± 0.237.3 ± 0.24.8 ± 0.1
Table 3. Result of the fit for the XPS peaks of the Pb(II) adsorbed nano-TiO2 samples.
Table 3. Result of the fit for the XPS peaks of the Pb(II) adsorbed nano-TiO2 samples.
SampleLevelBE (eV)at.%Bond TypeRef.
O 1s529.445.2O-Metal[37]
O 1s531.722.7O-adsorbed[36]
TiO2-Pb C2O 1s533.311.6OH-Organic[38]
1.16 mg L−1Ti 2p3/2458.118.4Ti+4 (TiOx)[35,36]
Pb 4f7/2138.12.1Pb+2[36]
O 1s529.446.6O-Metal[37]
TiO2-Pb C1O 1s531.517.4O-adsorbed[36]
47.54 mg L−1O 1s533.016.2OH-Organic[38]
Ti 2p3/2458.217.7Ti+4 (TiOx)[35,36]
Pb 4f7/2138.12.1Pb+2[36]
Table 5. Fit parameters of nonlinear models of adsorption kinetics of PFO, PSO, E, and IDM.
Table 5. Fit parameters of nonlinear models of adsorption kinetics of PFO, PSO, E, and IDM.
PFO ModelPSO ModelE Model
qe exp (mg g−1)47.15qe (mg g−1)46.29 (4)β (g mg−1)0.32 (4)
qe (mg g−1)46.28 (4)k2 (g mg−1 h−1)12.74 (5)α (mg h−1)43,223,954.34 (1)
k1 (h−1)16.29h (mg g−1 h−1)27,298.58
R20.990.990.92
RSS11.3811.38126.33
BIC5.906.0229.96
IDM
kp (mg g−1 h−0.5)5.04 (2)
C1 (mg g−1)26.28 (8)
R20.29
RSS1241.78
BIC52.82
Table 6. Parameters of the fitting of the adsorption isotherms of Langmuir, Freundlich, Temkin, Sips, and Redlich–Peterson models to the experimental data of the adsorption of Pb(II) on nano-TiO2.
Table 6. Parameters of the fitting of the adsorption isotherms of Langmuir, Freundlich, Temkin, Sips, and Redlich–Peterson models to the experimental data of the adsorption of Pb(II) on nano-TiO2.
LangmuirFreundlichTemkin
qe exp (mg g−1)47.5kF ((mg g−1)/(mg L−1)1/n)87.16 (1)KT (L g−1)1049.45 (7)
qm (mg g−1)65.99 (6)1/n0.24 (5)BT (J mol−1)219.85 (4)
kL (L mg−1)56.91 (2)n3.45 (7)
R20.840.690.83
RSS577.34577.341260.61
BIC48.3677.2556.95
SipsRedlich-Peterson
q m s (mg g−1)69.22 (1)A (L g−1)4191.41 (3)
k s   L m g 1 m s 50.71 (3)B (L mg−1)60.75 (3)
ms0.85 (5) β 0.97 (2)
R20.820.82
RSS570.48575.87
BIC48.2350.73
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vasquez-Caballero, M.A.; Canchanya-Huaman, Y.; Mayta-Armas, A.F.; Pomalaya-Velasco, J.; Checca-Huaman, N.-R.; Bendezú-Roca, Y.; Ramos-Guivar, J.A. Pb(II) Uptake from Polluted Irrigation Water Using Anatase TiO2 Nanoadsorbent. Molecules 2023, 28, 4596. https://doi.org/10.3390/molecules28124596

AMA Style

Vasquez-Caballero MA, Canchanya-Huaman Y, Mayta-Armas AF, Pomalaya-Velasco J, Checca-Huaman N-R, Bendezú-Roca Y, Ramos-Guivar JA. Pb(II) Uptake from Polluted Irrigation Water Using Anatase TiO2 Nanoadsorbent. Molecules. 2023; 28(12):4596. https://doi.org/10.3390/molecules28124596

Chicago/Turabian Style

Vasquez-Caballero, Miguel A., Yamerson Canchanya-Huaman, Angie F. Mayta-Armas, Jemina Pomalaya-Velasco, Noemi-Raquel Checca-Huaman, Yéssica Bendezú-Roca, and Juan A. Ramos-Guivar. 2023. "Pb(II) Uptake from Polluted Irrigation Water Using Anatase TiO2 Nanoadsorbent" Molecules 28, no. 12: 4596. https://doi.org/10.3390/molecules28124596

Article Metrics

Back to TopTop