Introduction

Arsenic (As) is amongst the highly abundant elements on the earth crust and known for its high toxicity. It exists in the ecosystem under diverse oxidation modes (− 3, 0, + 3 and + 5), but in water it commonly exists in the inorganic oxygenated form as arsenite (As (III)) as well as arsenate (As(V)), which are categorized as the most toxic species (Liu et al. 2022). Nowadays, the rapid development of industrial activities and civilization, continuous human activities along with the natural ecosystem actions have led to generation of multiple contaminates including arsenic species, and therefore, they dumped in the ecosystem from various implementations including fertilizers, pesticides, wastewater resulted from steel industry, smelting and volcanic eruptions (Zhu et al. 2021; Devi and Geethakarthi 2022). Subsequently, the consumption of water polluted with arsenic may cause severe poisoning effects such as blackfoot sickness, gastrointestinal diseases, keratosis, nerve damage and alteration in skin pigmentation and thickness in addition to several kinds of cancers such as bladder, skin, lungs and kidney (Susan et al. 2019). According to the WHO, the minimum permitted arsenic concentration in potable water must be ≤ 10 ppb (Babar et al. 2019). Hence, the elimination of arsenic from water is a priority task. This can be accomplished by some techniques as reverse osmosis, chemical sedimentation and ion exchange, but the majority of these techniques were faced with some disadvantages and limitations. For instance, the reverse osmosis membrane is an expensive technique and requires professional workers, while the chemical sedimentation technique causes direct interference by generation of other ions in water (Boussouga et al. 2021; Zeng et al. 2021; Tuzen et al. 2009; Uluozlu et al. 2010). Ion exchange technique is not a cost-effective process, besides many other restrictions (Chen et al. 2020). Adsorption, on the other hand, is more favorable than other remediation techniques due to facile fabrication of adsorbents, excellent processing, cheap and broad implementation domain (Kong et al. 2022; Ramalakshmi, et al. 2020). In addition, contaminated wastewaters with heavy metals have specific characteristics and behaviors, and therefore, the applied adsorbent should be selective for the target pollutant (Yang et al. 2022a, b; Sher et al. 2021). It has been illustrated that some adsorbents including carbon, carbon-supported composites, polymeric compounds, silicon-supported composites, metal oxides, oxy(hydro)oxides (iron, aluminum, titanium and manganese), and gels were utilized to eliminate arsenic species and other heavy metals from water (Sarker et al. 2017). On the other hand, these adsorbents have many limitations that prevented them from being widely applied in terms of facile diffusion, hard separation of the solid from liquid, and environmental risks (Feng et al. 2022).

Recently, great challenges in the field of nanotechnology have directed research efforts to fabricate efficient nanosorbents with some selected and aimed advantages as being cheap, reusable, with high elimination capacity, facile separation, and preserving the green ecological security (Amen et al. 2020). Between the widely utilized nanosorbents, LDHs represent a huge category of 2D interpolated composites as anionic clay compounds (Mahmoud et al. 2021a, b). Therefore, LDHs attracted great attentions by various scholars to apply in the wastewater remediation field (Mahmoud et al. 2022a, b). Additionally, LDHs showed very specific structures with positive metal hydrotalcite host layer via divalent metal cation (M2+) such as magnesium, ferrous or zinc along with a trivalent metal ion (M3+) as ferric, aluminum or chromium and interlayer anion An as carbonate, fluoride, nitrate, chloride or several organic anions which form opposite guest negative interlayer (Sun et al. 2022). The LDHs structure is generally expressed as [M1−× 2+ Mx3+(OH)2] [(An)x/n·mH2O], where (x) is the (M3+/(M2+ + M3+) with the molar percentage at which the net structure of LDHs is regularly determined by the following precise domain as 0.2 ≤ x ≤ 0.33 (Dinari et al. 2017). LDH properties are affected by several parameters as the kind and the percentage of divalent and trivalent cations. Also, it was reported that the valency of metal cations in LDHs layers determines its crystallinity, the interlayer anions type, the exterior sites and the charges of LDH layers, which efficiently affect their behavior (Wang et al. 2022). Consequently, several methods as coprecipitation, microwave-based technique, hydrothermal, ion interchange, solvothermal and other techniques were utilized to prepare LDHs (Janani et al. 2021). Also, the LDH materials are characterized with several specifications including elevated surface area, interlinkage of positive metal cation exterior with contaminants, hydrophilicity, high interlayer negative charge density, controllable interlayer negative ions, adjustable grain dimension, low cost, ecofriendly and ability to remove pollutants from water with high capacity (Chengqian et al. 2023; Yuan et al. 2022; Feng et al. 2022). Moreover, the advantages of LDHs have distinguished them with superior anion-exchange characteristics in positive laminates and interlayer void, so LDHs were identified as excellent candidates and efficient adsorbents for water purification and elimination of several pollutants including inorganic anions, organic anions and oxygen containing anions (Mahmoud et al. 2022a, b). As previously reported, bimetallic LDHs as magnesium ferric, cobalt ferric, magnesium aluminum and calcium aluminum were found to be efficient adsorbents for elimination of As(V) from water matrices (Choong et al. 2021). Also, it was reported that modified ferric and nickel LDH with diatomite was efficient for As(V) and As(III) adsorption, while zinc ferric LDH was used for arsenite elimination (Dai et al. 2022; Wang et al. 2021). In addition, functionalized magnesium and aluminum LDH with magnetic hollow Fe3O4 was reported for arsenate elimination (Mahmoud et al. 2022a, b). Regrettably, conventional LDHs have many drawbacks as limited active sits or structural defects as the ultra-thin thickness structure making them difficult to be reused or detached from water posterior adsorption, which limited the experimental applications of LDHs (Feng et al. 2022). Also, some LDHs such as nickel ferric LDH have shown some sort of agglomeration, which decreased their surface area and active sites (Wang et al. 2021). So, the LDHs should be modified to increase their active sites via interjection, exterior modification and stuffing, aiming to generate selective, reusable and highly efficient LDHs (Feng et al. 2022). Furthermore, the LDH materials could be modified with compounds like carbonaceous nanomaterials, negative ions, surfactants, iron nanocomposites and magnetic carriers to induce LDHs with unique surface and excellent adsorption specifications (Tao et al. 2017).

Many cost-effective and efficient nanocomposites were investigated for enhancing the properties of LDHs (Wang et al. 2022). Moreover, two-dimensional (2D) structures and composites have recently grabbed great attention owing to their magnificent advantages compared to their peers (Weng et al. 2023). Therefore, LDH materials were aimed to modify in order to synthesize adsorbents with some incorporated specifications via integration of the distinctive original composite characteristics (Zheng et al. 2023a, b). Generally, the 2D structures could be classified into metal dichalcogenides, metal carbides or nitrides (MXenes), organic compounds as covalent organic frameworks and metal organic frameworks and Xenes (Torres et al. 2022). Xenes were defined and produced as a category of layered formula composites by incorporation of single layer of one element (Baig 2023). Additionally, Xenes have prominent physical specifications with wide applications in several scopes because of their elevated atomic utilization, high surface area, unique crystal shape and excellent stability (Zhang et al. 2022). However, Xenes are a modernly protruded category of 2D nanostructures, so their synthesis and application history are still recent and their applications are mainly limited to materials science and not widely applied in water purification (Huang et al. 2020). Bismuthene is an example of 2D nanostructure and known to exhibit unique specifications such as elevated electron motions, increased specific capacity, less toxicity, insulation properties and great magneto-impedance influence (Zhang et al. 2021). Therefore, it was utilized in several fields as high-speed photonics, thermoelectrical applications, spintronic instruments, non-reduced temperature quantum spin Hall compounds, energy preservation in sodium-ion batteries, heavy metals identification and electrocatalysis (Zhang et al. 2022). So, as previously reported, bismuthene-based compounds have shown specific physicochemical advantages as being non-carcinogenic and nontoxic and shown promising applications in the elimination of anionic inorganic contaminates as phosphate, nitrate, fluoride, sulfate and chloride (Balint et al. 2021; Saddique et al. 2023). Moreover, bismuth hydroxide was previously applied for arsenic capturing from water (Franceschini et al. 2022).

So far, the application of 2D-bismuthene or modified 2D-bismuthene composites for water remediation from As(V) has not been previously reported. Therefore, the novelty in this study in mainly focused and aimed to synthesize 2D-bismuthene (Biene) embedded into zinc aluminum bismuth-layered double hydroxides to obtain unique incorporated specifications of the assembled ZnAlBi LDHs-embedded-Biene as a pioneer structure with hybrid nanosheet system due to the combination of metal oxyhydroxide, metal oxides, bismuthene and carbonaceous anion in the inner layer. The synthesized ZnAlBi LDH-embedded-Biene with multiple layers, crystalline, mesoporous and ecofriendly characters is expected to be a stable nanosorbent in water treatment and elimination with high efficiency for As(V) removal.

Experimental

Chemicals and solutions

Bi(NO3)3·5H2O (FW 485.07—Assay 98.5%), Zn(NO3)2·6H2O (FW 297.47—Assay 96%), Al(NO3) 39H2O (FW 375.13—Assay 98.0–102.0%), polyvinylpyrrolidone (PVP) (FW 58,000 g/mol), disodium hydrogen arsenate (Na2HAsO4, FW 312.01—Assay 98%), ammonium molybdate (NH4)2MoO4 (FW 200.06 g/ml—Assay 99.98%) and l-ascorbic acid (C6H8O6, FW 176.124, Assay 30%) were purchased from Lobachemie, India. Na2CO3 (FW 105.9888—Assay 99.88%), NaOH (FW 40—Assay 98%), ethylene glycol (C2H4(OH)2, FW 62.068—Assay 80%), hydrazine monohydrate (N2H4.H2O, FW 50.06—Assay ≥ 98), NaCl (FW 58.44,—Assay 100%), NH3.H2O (FW 35.046—Assay 30%), HNO3 (FW 63.013—Assay 99.9%) and HCl (FW 36.46,—Assay 37%), were obtained from VWR international Ltd, England.

Instrumentations

All applied instruments are listed in Table 1.

Table 1 Various instrumentations

Synthesis of Bismuthene (Biene)

Bismuthene (Biene) was synthesized by bottom-up hydrothermal process (Zhang et al. 2021). Briefly, 0.3 g of PVP was added to 36 mL of (EG) and agitated for 15 min and 2.8 g of Bi(NO3)3·5H2O was then added, followed by adding 12 mL hydrazine and 72 mL aqueous ammonia. The resulted mixture was stirred for 1 h, autoclaved and heated at 100 °C for 12 h. Finally, the autoclave was left to cool, and the resulted Biene was centrifuged, collected, washed by ethanol and dried in a furnace at 60 °C as sketched in Scheme 1.

Scheme 1
scheme 1

Synthesis of Biene and ZnAlBi LDHs-embedded-Biene nanosorbent

Synthesis of embedded bismuthene into zinc aluminum bismuth LDHs (ZnAlBi LDHs-embedded-Biene)

ZnALBi LDHs were synthesized via coprecipitation process followed by a hydrothermal treatment method (Dinari et al. 2017) in which bismuthene (Biene) was added to (ZnALBi LDHs) with a ratio of 1:2, respectively. In 100 mL DW, 1.3 g of Biene was stirred for 15 min in a mixture of 3.0 g of Zn(NO3)2·6H2O and 1.41g Al(NO3) 39H2O until all salts were dissolved (solution A). 0.75 g Bi(NO3)3·5H2O was added in 5 mL HNO3 (solution B). Then, the two solutions were mixed and adjusted at pH = 9 via mixture of 2.0 M NaOH and 0.5 M Na2CO3 and stirred for 12 h. Thence, this was autoclaved and heated at 160 °C for 24 h and, finally, centrifuged and washed several times by distilled water till neutralization. The resulted solid (ZnAlBi LDHs-embedded-Biene) was dried in oven at 60 °C for 12 h as sketched in Scheme 1.

Uptake studies of As(V) by ZnAlBi LDHs-embedded-Biene

A stock solution of disodium hydrogen arsenate (1000 mg L−1) was prepared from which 5–25 mg L−1 concentrations were obtained. As(V) was detected via a UV–visible 2700 spectrophotometer instrument at λmax = 840 nm (Hou et al. 2016) where the arsenate was detected by mixing 2 mL solution with 0.4 mL of 6.5 gL−1 (NH4)2MoO4) and 0.2 mL of 100 gL−1 L-ascorbic acid. This was diluted with 10 mL DW and left for 90 min, and finally, the absorbance was measured. The uptake reaction of As(V) from water was done at room temperature, and the mechanical shaking was accomplished by an automatic shaker. The As(V) uptake behavior from water was investigated under various variables such as (ZnAlBi LDHs-embedded-Biene) mass, As(V) concentration, reaction time, solution pH, ionic strength and temperature.

PZC of ZnAlBi LDHs-embedded-Biene and impact of pH on the uptake of As(V)

The pH of As(V) concentrations (5–25 mg L−1) was adjusted at pH 2 to 11 via 0.1 M NaOH and 0.1 M HCl. 20 mL of As(V) with various pHs was added to 10 mg of ZnAlBi LDHs-embedded-Biene and vibrated for 30 min. Finally, after the separation of the unreacted As(V) by filtration, the remaining As(V) in the filtrate was detected as stated (Sect. "Uptake studies of As(V) by ZnAlBi LDHs-embedded-Biene").

To measure the PZC of ZnAlBi LDHs-embedded-Biene, a 0.1 g of nanosorbent was mixed with 40 mL of 0.01 potassium chloride, and the solution pHs were adjusted in the range from 2 to 11 by 0.1 M NaOH or 0.1 M HCl and agitated for 4 h. Then, the final pHs of all solutions were detected after 12 h. The PZC of ZnAlBi LDHs-Embedded-Biene was finally detected and identified by plotting ΔpH versus pH.

Impact of ZnAlBi LDHs-embedded-Biene mass on the uptake of As(V)

2 mg to 50 mg of ZnAlBi LDHs-embedded-Biene was added to 20 mL of 5–25 mg L−1 As(V) solutions at pH 3 and shaken for 30 min, and after separation, the absorbance of solution was detected at λmax = 840 nm as stated above.

Impact of ionic strength on the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

10 mg of ZnAlBi LDHs-Embedded-Biene was added to 20 mL of 5–25 mg L−1As(V) concentrations at pH 3; then, NaCl (10 to 100 mg) was added and shaken for 30 min. After nanosorbent separation, the absorbance was detected as stated above.

Impact of interaction duration on the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

10 mg of ZnAlBi LDHs-Embedded-Biene was added to 20 mL of 5–25 mg L−1 As(V) concentrations at pH 3 and shaken at different interaction duration intervals from 2 to 60 min, and after separation by filtration, the absorbance of retained As(V) concentrations was detected as stated above.

Impact of As(V) concentration on the uptake by ZnAlBi LDHs-embedded-Biene

20 mL of different concentrations of As(V) concentrations from 5 to 50 mg L−1 was adjusted at pH 3 and then added to 5 mg, 10 mg and 20 mg of ZnAlBi LDHs-embedded-Biene. These solutions were shaken for 30 min, and after separation, the absorbance of retained As(V) concentrations was detected as stated above.

Impact of temperature on the adsorptive uptake of As(V) by ZnAlBi LDHs-embedded-Biene

20 mL of 5–25 mg L−1As(V) concentrations at pH 3 was added to 10 mg of ZnAlBi LDHs-Embedded-Biene, and the uptake temperatures of solutions were controlled by a thermostat in the range from 25 to 60 °C followed by shaking for 30 min, and after separation, the absorbance of retained As(V) concentrations was detected as stated above.

Application of ZnAlBi LDHs-embedded-Biene in adsorptive uptake of As(V) from water samples

As(V), 5–15 mg L−1, solutions were prepared by spiking into sea water, tap water in addition to wastewater specimens, and the pH of these samples was adjusted at pH 3. Twenty milliliters of each sample was added to 10 mg of ZnAlBi LDHs-Embedded-Biene, shaken for 30 min, and after separation, the absorbance of retained As(V) was detected as stated above.

Regeneration and recyclability of ZnAlBi LDHs-embedded-Biene

250 mg of ZnAlBi LDHs-embedded-Biene was added to 40 mL of 100 mg L−1 As(V) and shaken 30 min. The separated was washed with 100 mL 0.1 mol L−1 HCl—DW to neutralization and finally dried in furnace at 60 °C. Hence, 10 mg was treated with 20 mL of 5 mg L−1 As(V) adjusted at pH 3 and shaken for 30 min. After separation, the absorbance of retained As(V) concentrations was detected as stated above.

Results and discussion

Characterization of ZnAlBi LDHs-embedded-Biene

FTIR of Biene is illustrated in Fig. 1a and found to exhibit a peak at 3421 cm− 1 assigned to hydroxyl bond on the bismuthene exterior, while the bands at 1657 cm− 1 and 1386 cm− 1 are assigned to the carbonyl group of the bismuth sheets on the surface of bismuthene nanosheets (Girirajan et al. 2022). ZnALBi LDHs spectrum shows two bands at 1633 cm−1 and 3428 cm−1 for water hydroxyl groups on the exterior and interior surfaces of ZnALBi LDHs, while the band at 1363 cm−1 is referring to O–C–O vibration due to the incorporation of CO32−and interlayer water via powerful hydrogen bond (Dinari et al. 2017). Additionally, the metal oxygen bonds in the form of tetrahedral and octahedral stretching modes are detected at 967 cm−1, 910 cm−1 and 847 cm−1 for O–Al–O bond, while the band at 610 cm−1 is referring to Al-O. The 789 cm−1 and 553 cm−1 peaks confirm the O–Zn–O and Zn–O vibration modes, and the characteristic band at 431 cm−1 is confirming the Bi–O bond (Jatav et al. 2022; Adeniyi et al. 2022). The FTIR of ZnAlBi LDHs-embedded-Biene shows two bands at 3403 cm−1 and 1639 cm−1 for hydroxyl group and water molecule δ(H–O–H) interlayer, while the 1390 cm−1 band refers to the presence of CO32− in the inner layer of bismuth sheets (Girirajan et al. 2022). The same stretching vibrations for O–Al–O bond are detected by the peaks at 998 cm−1 and 844 cm−1, while those related to O–Zn–O and Zn–O bonds stretching are evident by two bands at 664 cm−1 and 545 cm−1. Moreover, an appearing band at 452 cm−1 is related to the Bi–O bond stretching mode (Feng et al. 2021). Finally, the FT-IR spectrum of loaded ZnAlBi LDHs-embedded-Biene with As(V) is characterized by low intensity of a peak at 1390 cm−1 due to interjection of As(V) into the inner layer via an anion exchange and inner-layer anions (Solís–Rodríguez et al. 2021), while the related bands to Zn–O, Bi–O, O–Zn–O and O–Al–O bonds were disappeared with the appearance of a band at 846 cm−1 for the generated As-O bond (Jung et al. 2021a, b).

Fig. 1
figure 1figure 1

a. FT-IR spectra of Biene, ZnALBi LDHs, ZnAlBi LDHs-embedded-Biene and ZnAlBi LDHs-embedded-Biene with loaded As(V), b. XRD Patterns of ZnAlBi LDHs-embedded-Biene nanosorbent, c. The XPS of ZnAlBi LDHs-embedded-Biene nanosorbent, d. TEM image of ZnAlBi LDHs-embedded-Biene nanosorbent, e. SEM image of ZnAlBi LDHs-embedded-Biene nanosorbent, f. The EDX of ZnAlBi LDHs-embedded-Biene nanosorbent, g. Adsorption–desorption isotherms of ZnAlBi LDHs-embedded-Biene nanosorbent, h. TGA thermogram of ZnAlBi LDHs-embedded-Biene nanosorbent

As shown in Fig. 1b, the XRD of ZnAlBi LDHs-embedded-Biene revealed diffraction peaks at 2θ 11.5°, 23.8°, 30° and 32.7°, which assigned to (003), (006) and (012) planes indicating the crystalline and layered formula similar to the hydrotalcite compounds (Dinari et al. 2017; Wan et al. 2019; Kumari et al. 2021). The appeared bands at 30°, 42.2°, 47.5°, and 56.8° are corresponding to the lattice planes (012), (110), (202) and (024) for Biene nanolayer with rhombohedral formula as the layers are joined via van der Waals force amidst the inner sheets (Girirajan et al. 2022; Xu et al. 2022; Wei et al. 2022). Therefore, the XRD of ZnAlBi LDHs-embedded-Biene exhibited good crystalline structure and assured the successful modification process.

The XPS is utilized to elucidate the element assembly as well as the exterior valence status of the fabricated materials. The XPS survey (Fig. 1c) revealed the existence of elements Bi, Zn, O, C, besides Al with percentages 9.55%, 7.14%, 41.27%, 38.39% and %3.65, respectively. As clarified in Fig. 1c, the Bi 4f is divided into two bands at 158.78 eV and 164.06 eV conformable to Bi-4f7/2 and Bi-4f5/2, respectively (Zhang et al. 2021; Girirajan et al. 2022). The bands at 159.73 eV and 165.25 indicate the oxidation state of Bi, and the 160.02 eV band refers to the presence of metallic bismuth (Torres et al. 2022; Zhang et al. 2021, 2022). Furthermore, the bands at 1021.8 eV and 1044.8 eV are congruent with Zn 2p3/2 and Zn 2p1/2 signals, respectively, to outline the regular Zn2+ bands in ZnALBi LDHs. The 74.1 eV band is assigned to 2p1/2 signals to confirm the existence of aluminum oxyhydroxide (Al–OH) bonds (Kumari et al. 2021; Santamaría et al. 2020). The O1s band at 529.83 eV indicates the existence of metal oxide (M−O) (Vasseghian et al. 2023), while the 531.79 eV band is related to metal hydroxide (M–OH) bonds and also to the occurrence of C=O (Kumari et al. 2021; Meng et al. 2023). The C 1s band at 285.54 eV elucidates the existence of C–O link to clarify the occurrence of CO32− in the inner laminate of ZnALBi LDHs (Haddadi et al. 2022; Zheng et al. 2023a, b).

The morphology structure and particle size of ZnAlBi LDHs-embedded-Biene were determined by TEM, SEM, and EDX as represented in Fig. 1d. The TEM image refers to the occurrence of globular particles that are arranged and embedded in the formed sheets with homogeneous distribution to confirm their ultra-thin composition and symmetric thickness. The detected average particle size in the domain was identified in the range 5–8 nm, while the illustrated SEM image in Fig. 1e confirms the layered formula of ZnAlBi LDHs-embedded-Biene (Dinari et al. 2017; Adeniyi et al. 2022). The EDX analysis of ZnAlBi LDHs-embedded-Biene refers to the elemental composition as shown in Fig. 1f. The existence of Bi, O, Zn, Al and C constituents can be elucidated to certify the regular apportionment of Bi and O across the sheets due to the presence of bismuthene and the successful integration between bismuthene and ZnALBi LDHs (Dinari et al. 2017; Girirajan et al. 2022).

As illustrated in Fig. 1g, the N2-isotherm of ZnAlBi LDHs-embedded-Biene refers to type IV isotherms with H3 hysteresis loop associated with silt form pores of packing lamina like particles to confirm the physical adsorption of N2 according to the layered structure. The specific surface area was found 20.03 m2 g−1, while the pore size and its volume were detected by BJH method as 47.721 nm and 0.2389 cm3 g−1, respectively. So, the fabricated ZnAlBi LDHs-embedded-Biene could be described as a mesoporous structure (Kumari et al. 2021; Liu et al. 2018).

TGA is utilized to examine the thermal stability of ZnAlBi LDHs-embedded-Biene in temperature ranging from 30 to 700 °C as illustrated in Fig. 1h. Primarily, the temperature from 29.5 to 151.3 °C shows mass loss of 0.063 mg with percentage 1.12%, while at the second plateau of temperature from 151.3 to 184.6 °C exhibits mass loss of 0.080 mg with percentage 1.42%, which are mainly produced by surface water evaporation as well as the inner-layered water, respectively. Furthermore, at the third the temperature ranging from 184.6 to 377.4 °C, the mass loss is 0.279 mg with percentage 4.97% that is attributed to dehydroxylation and decarbonation. The fourth step from 377.4 to 453.3 °C with mass loss of 0.207 mg and percentage of 3.69%. The last step from 453.3 to 552.9 °C with mass loss of 0.091 mg and percentage of 1.62% is due to degradation of the mixed metal oxide of (ZnALBi LDHs) along with the dissociation of CO32− anion from inner layer via emission of carbon dioxide gas (Yang et al. 2022a, b; Kang and Ye 2022).

Batch uptake studies of As(V) by ZnAlBi LDHs-embedded-Biene

The uptake of As(V) from water by ZnAlBi LDHs-embedded-Biene was investigated by the batch uptake operation based on determination of the uptake percentage of As(V) as stated in Eq. (1).

$$\mathrm{\% Uptake }=\frac{{{\text{C}}}_{{\text{o}}}-{{\text{C}}}_{{\text{e}}}}{{{\text{C}}}_{{\text{o}}}} \times 100$$
(1)

While the As(V) equilibrium uptake capacity (qe) can be identified by Eq. (2).

$${q}_{e}=\frac{{C}_{o}-{C}_{{\text{e}}}}{W/V}$$
(2)

where Co and Ce are the initial and equilibrium As(V) concentrations, while V is the volume (L) and W is the mass of ZnAlBi LDHs-embedded-Biene in grams (Mahmoud et al. 2020).

PZC of ZnAlBi LDHs-embedded-Biene and impact of pH on the AS(V) uptake

PZC is defined as an important experiment to detect the ZnAlBi LDHs-embedded-Biene surface charge (Yu et al. 2021). It is mainly based on detecting the exterior charge at several pH values via combining the nanosorbent with solutions adjusted at several pH values under fixed solution ionic strength. The PZC of ZnAlBi LDHs-embedded-Biene was found to be at pH 8 as illustrated in Fig. 2a. Thus, at pH < 8, the exterior surface charge was mainly positive, and at pH > 8 the exterior surface charge is mainly negative (Simsek et al. 2022).

Fig. 2
figure 2

a. The PZC of ZnAlBi LDHs-embedded-Biene, b. Impact of pH on the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

The ZnAlBi LDHs-embedded-Biene uptake efficiency was explored using pH ranging from 2 to 11. It can be observed that the ultimate uptake efficiency was fulfilled at pH 3 with uptake percentages 94.67%, 90.74% and 84.82% for 5, 10 and 25 mg L−1As(V) concentrations, respectively. The uptake efficiencies were gradually decreased with increasing the pH till reaching the lowest values at pH 11 with percentages 74.33%, 51.43% and 36.84% for the same three concentrations, respectively, as illustrated in Fig. 2b. Generally, As(V) ion exists in the form of H2AsO4−1 structure at pH from 2.1 to 6.7, and when the pH value exceeds 6.7, the predominant species is HAsO4−2. However, the PZC of ZnAlBi LDHs-embedded-Biene was detected at pH 8, and this means that the exterior surface is positive at pH < 8. This explains the high removal efficiency at pH 3 due to the direct electrostatic interaction between negative As(V) with positively charged exterior surface (Mahmoud et al. 2022a, b). In addition, the inner interchangeable anions as CO32− and metal hydroxides in semi-hydrotalcite structure of the nanosorbent show exchangeable behavior with As(V) ions, which enhanced the uptake efficiency by ZnAlBi LDHs-embedded-Biene via anion exchange mechanism. The decline in the uptake of As(V) with increasing the pH can be explained by the repulsion between the negatively charged ZnAlBi LDHs-embedded-Biene exterior surface and the As(V) at pH > 8. Moreover, with increasing the pH, OH ions concentration increased and competed with As(V) ions (Nie et al. 2022).

Impact of ZnAlBi LDHs-embedded-Biene mass on the As(V) uptake

The As(V) uptake by ZnAlBi LDHs-embedded-Biene was detected by using mass ranging from 2 to 50 mg. It was noticed that the uptake capacities increased from 54.25%, 47.32% and 32.43% using 2 mg of nanosorbent to 99.46%, 98.19% and 95.29% for concentrations 5, 10 and 25 mg L−1, respectively, by using 50 mg of nanosorbent as illustrated in Fig. 3. This behavior can be explained on the basis of increasing active sites on the nanosorbent exterior surface and the inner exchangeable anions by increasing the adsorbent mass (Nie et al. 2022).

Fig. 3
figure 3

Impact of ZnAlBi LDHs-embedded-Biene mass on the uptake of As(V)

Ionic strength impact on As(V) uptake

The presence of salts in the reaction medium was found to influence and control the uptake capacity of adsorbate by adsorbent (Zhang et al. 2017). The As(V) uptake by ZnAlBi LDHs-embedded-Biene was detected by using various masses of sodium chloride in the range from 10 to 100 mg. It was found that the uptake capacity of As(V) was slightly influenced by the salt, and it was found that the uptake percentages were 93.4%, 90.11% and 83.82% for 5, 10 and 25 mg L−1 As(V), respectively, upon using 10 mg salt dose, while the uptake percentages were found to be 77.74%,74.4 1%, 68.71% upon using 100 mg salt dose as illustrated in Fig. 4. Thus, the As(V) uptake was declined by about 15% to indicate the good selectivity of ZnAlBi LDHs-embedded-Biene for As(V). Furthermore, the uptake of As(V) ions through the exterior sphere complexation was influenced by the increase in ionic strength as the uptake process was directly influenced by the electrostatic force between adsorbent and adsorbate, which decreased due to the existing ions as Na+ and Cl. On the other hand, the uptake by the interior sphere coordination was not affected by the ionic strength as the interaction, in this case, is based on an ion ligand exchange and the powerful chemical bond formation. Therefore, the proposed uptake mechanism of As(V) by ZnAlBi LDHs-embedded-Biene is mainly based on both the exterior sphere complexation along with the interior sphere coordination (Liu et al. 2019).

Fig. 4
figure 4

The impact of ionic strength on the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

Impact of interaction duration on the As(V) uptake and kinetic studies

The uptake behavior of As(V) ions onto ZnAlBi LDHs-embedded-Biene was investigated by using time range from 2 to 60 min. The uptake process was rapidly elevated at the first 25 min of reaction due to the availability of huge number of active sites on the surface of ZnAlBi LDHs-embedded-Biene along with the increased number of interchangeable CO32− anions in the inner layer till reaching equilibrium at 30 min at which all the sites and interchangeable anions were loaded with As(V). As illustrated in Fig. 5a, the uptake capacity was elevated from 36.17%, 31.18% and 26.06% at the first 2 min to reach to 94.27%, 90.87% and 85.64% at 30 min by using 5, 10 and 25 mg L−1 As(V), respectively (Mahmoud et al. 2022a, b).

Fig. 5
figure 5

a. The impact of interaction duration on the uptake of arsenate by ZnAlBi LDHs-embedded-Biene, b. Pseudo-first order model of the uptake reaction of Arsenate by ZnAlBi LDHs-embedded-Biene, c. Pseudo-second order model of the uptake reaction of Arsenate by ZnAlBi LDHs-embedded-Biene

For more investigation of the uptake mechanism of As(V) by ZnAlBi LDHs-embedded-Biene, several kinetic models were employed as pseudo-first order, pseudo-second order, intraparticle diffusion and Elovich for detection of the ultimate convenient model. The uptake kinetics models can be described as an essential manner to detect the uptake optimum time in addition to determine the concentration of adsorbate, which is transferred to the solid sorbent and the remaining adsorbate in the solution (Simsek et al. 2022). Firstly, the pseudo-first order (Wang et al. 2020) besides pseudo-second order (Kasiuliene et al. 2020) models were explored with their equations stated (Table 2). The computed R2 from pseudo-first order were 0.9688, 0.9416 and 0.9412 for 5, 10 and 25 mg L−1 of As(V), while the pseudo-second order showed R2 = 0.9889, 0.9848 and 0.9821 as stated in Table 3 and illustrated in Fig. 5b and c, respectively. Hence, pseudo-second order established the best model for As(V) elimination reaction more than pseudo-first order. Moreover, the (qecal) was found very close to (qeexp) in pseudo-second order model rather than in pseudo-first order model as stated in Table 3 to confirm that the average uptake was dependent on the number of sites on ZnAlBi LDHs-embedded-Biene (Huo et al. 2022). As also stated in Table 3, k2 was found to decline from 0.013118 (g/mg min) for 5 mg L−1 As(V) to 2.1134 × 10–3 (g/mg min) for 25 mg L−1 As(V) concentration to reveal that upon decreasing the concentration of As(V), ZnAlBi LDHs-embedded-Biene was characterized by rapid elimination of As(V) from water (Wang et al. 2023).

Table 2 Various kinetic models for As(V) uptake by ZnAlBi LDHs-embedded-Biene
Table 3 Computed kinetic data for As(V) uptake by ZnAlBi LDHs-embedded-Biene

The intraparticle diffusion expression was also applied as the uptake reaction of As(V) depends on two essential stages. Primarily, the As(V) diffusion into the adsorbent exterior positions followed by the second spreading step into the pores of ZnAlBi LDHs-embedded-Biene (Bujdak 2020). The mathematical calculation of this model is stated in Table 2. The values of R2 were 0.906, 0.896 and 0.8974 for 5, 10 and 25 mg L−1 As(V), while kid and C are stated in Table 3. It was also concluded that the straight lines were passing by the origin to indicate that the intraparticle diffusion was not the main stage to describe the As(V) uptake reaction by ZnAlBi LDHs-embedded-Biene. The graph indicated that the As(V) uptake reaction was correlated to three stages, rapid uptake followed by the second stage at which stable dominating uptake process of intraparticle diffusion and finally, the last stage at which equilibrium was established (Dudek and Kołodyńska 2022). It was observed that the lower As(V) concentration (5 mg L−1) was more fitting to this intraparticle diffusion model than higher ones according to (R2).

Elovich is last evaluated expression to figure out the heterogeneous surface as a convenient model for chemisorption kinetics (Giri et al. 2021) as listed in Table 2. The R2 were 0.9352, 0.9276 and 0.9287 for 5, 10 plus 25 mg L−1 As (V), while (β) and (α) were also calculated. The elevating (α) from 4.2969 (mg/g min) at 5 mg L−1 As(V) to 12.9308 (mg/g min) at 25 mg L−1 As(V) refer to a chemisorption behavior to elucidated the occurrence of ion exchange between the inner carbonate anion and arsenate (Ahmad et al. 2020). In addition, the decrease of (β) from 0.4828 to 0.09599 (mg /g) is corresponding to the exterior coverage that tend to decline the availability of reactive sites (Dudek and Kołodyńska 2022).

Based on R2 values as well as other correlated parameters, it can be concluded that the uptake mechanism of As(V) by ZnAlBi LDHs-embedded-Biene was highly described by pseudo-second-order model.

Impact of concentration on As(V) uptake by ZnAlBi LDHs-embedded-Biene and adsorption isotherms

The ranging concentration from 5 to 50 mg L−1 As(V) was investigated in presence of various masses of ZnAlBi LDHs-embedded-Biene (5–20 mg) to explore the impact of As(V) initial concentration on the efficiency of this nanosorbent. As shown in Fig. 6a, it is clear that upon elevating the As(V) concentration, the elimination percentages were decreased from 94.83%, 98.84% and 90.017% (5 mg L−1 As(V)) to 57.21%, 62.67% and 52.68% at (50 mg L−1 As(V)) using 5 mg, 10 mg plus 20 mg ZnAlBi LDHs-embedded-Biene, correspondingly, while the uptake values (qe mg g−1) were elevated with increasing the As(V) concentration (Jung et al. 2021a, b). Thus, the decline in uptake percentage values can be explained by the decrease in the number of active sites on the nanosorbent surface. Additionally, the existing exchangeable carbonate anions in the inner layer of nanosorbent became fully replaced by As(V). The increase in the As(V) uptake (qe) with elevating the As(V) concentration can be explained by the increasing in the As(V) moving force, which make their spreading on the nanosorbent to occur rapidly and predominate the As(V) mass transferring impedance from solution to the nanosorbent (Lee et al. 2019; Mahmoud et al. 2021a, b).

Fig. 6
figure 6

a. Impact of initial As(V) concentration on the uptake values by ZnAlBi LDHs-embedded-Biene, b. Langmuir isotherm for the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

Adsorption isotherms were also tested and evaluated to detect the possible interaction mechanisms between As(V) and ZnAlBi LDHs-embedded-Biene by checking various adsorption isotherms as Langmuir, Freundlich, Dubinin–Radushkevich (D-R), Redlich–Peterson in addition to Temkin. Initially, Langmuir model was checked as it supposes that the As(V) uptake reaction is based on a mono-layer process on a homogeneous exterior surface without interaction among the uptaken species. As represented in Table 4 and Fig. 6b, the R2 = 0.9931, 0.9867 and 0.9795 by using 5 mg, 10 mg plus 20 mg ZnAlBi LDHs-embedded-Biene, correspondingly of ZnAlBi LDHs-embedded-Biene, respectively. The Langmuir separation factor (RL) values were (0.43396–0.0712), (0.3823–0.0583) and (0.28164–0.03773) using 5 mg, 10 mg plus 20 mg doses of ZnAlBi LDHs-embedded-Biene. All these values indicate favorable uptake reactions of As(V) by ZnAlBi LDHs-embedded-Biene. It is known that when 0 < RL < 1, it means a favorable operation; if RL = 0, it confirms an irreversible operation, if RL > 1, it denotes to an unfavorable operation, and if RL = 1, it refers to a linear operation. Furthermore, the qmax was computed as 119.047 (mg/g), 63.29 (mg/g) and 33.113 (mg/g) using 5 mg, 10 mg plus 20 mg nanosorbent correspondingly, and these values are relatively higher than other listed adsorbents in Table 5 to emphasize the efficiency of ZnAlBi LDHs-embedded-Biene nanosorbent (Kanagavalli et al. 2021).

Table 4 Examined isotherm models for As(V) uptake by ZnAlBi LDHs-embedded-Biene
Table 5 Comparisons of As(V) uptake (qmax) by ZnAlBi LDHs-embedded-Biene and other adsorbents

The Freundlich adsorption isotherm (Wang and Guo 2020) assumed that multiple layer uptake interactions can take place on a heterogeneous exterior surface, which includes asymmetrical obtainable positions with alternative uptake energy. As stated in Table 4, R2 were 0.9769, 0.9928 and 0.994 by using 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly. The n-values were 2.1767, 2.4444 and 3.123 by using 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly. This is conformable to convenient uptake process of As(V) by ZnAlBi LDHs-embedded-Biene. If this value is from 2 to 10, it indicates a convenient uptake operation, while if 1 < n < 2 it means a moderate uptake operation and if n is less than one, it identifies a difficult uptake process (Mallakpour et al. 2023). Additionally, the elevated Kf 28.380 (L. mg−1) by 5 mg of nanosorbent refers to high uptake capacity (Pathania et al. 2022).

Temkin isotherm (Shikuku et al. 2018; Lou et al. 2022) assumed that the uptake heat of adsorbate particles (As(V)) directly decreases with increasing the nanosorbent exterior sites owing to the combination with As(V) ions as the model is adequate to describe the electrostatic interaction based on chemical adsorption operation. As mentioned in Table 4, the R2 were 0.9867, 0.9674, 0.9178 by 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly along with the other computed parameters as outlined in Table 6. The Redlich–Peterson isotherm is a multilateral model of mixed experimental isotherm models based on the integration between Freundlich and Langmuir. It denotes to homogenous and heterogeneous systems according to the listed mathematical statement in Table 4. The characterized R2 values from this model were 0.9835, 0.9965 and 0.9992 using 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly. The g and aRP were also calculated and are listed out in Table 6. The (g) exponential domain is (0 ≤ g ≤ 1) for Freundlich, and if g = 1, this model is transformed to Langmuir model, while if g is zero, the model is correlated to linear expression (Pathania et al. 2022).

Table 6 Computed isotherm parameters for As(V) uptake by ZnAlBi LDHs-embedded-Biene

The D-R model was also utilized for the estimation of As(V) uptake by ZnAlBi LDHs-embedded-Biene as well as the porosity and the energy of uptake. The mathematical equation is outlined in Table 4, which also compiles R2 = 0.8052, 0.7114 and 0.7065 using 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly to confirm the invalidity of this model to interpret the data. In addition, Es were found as 1.58, 7.07, 4.0824 (kJ/mol) by 5 mg, 10 mg plus 20 mg nanosorbent, correspondingly to express the uptake operation of As(V) by ZnAlBi LDHs-embedded-Biene as mainly a physisorption operation (Tsamo et al. 2017; Mallakpour et al. 2022).

It is evident from the concluded data that the related models to Langmuir, Freundlich and Temkin isotherms showed high appropriateness according to R2 values along with RL,, n and other mentioned parameters to confirm that these three models can be applied to identify the uptake operation of As(V) by ZnAlBi LDHs-embedded-Biene.

Impact of temperature on As(V) uptake by ZnAlBi LDHs-embedded-Biene and thermodynamic parameters

The As(V) uptake by ZnAlBi LDHs-embedded-Biene at diverse temperature was studied in the temperature range from 25 to 60 °C. As represented in Fig. 7a, the uptake percentages of As(V) declined from 93.24%, %90.38 and 84.299% to 70.33%, 65.22% and 54.63% upon using 5, 10 and 25 mg L−1 As(V) versus temperature elevation. Also, the thermodynamic parameters (ΔG°), (ΔS°) and (ΔH°) were calculated from the Van’t Hoff plot as mentioned in Eqs. (35).

$$\ln \, K_{D } = \Delta S^\circ /R - \Delta H^\circ /RT$$
(3)
$$\Delta G^\circ = - RT \, \ln K_{D }$$
(4)
$$K_{D } = q_{e} /c_{e}$$
(5)
Fig. 7
figure 7

a. Impact of temperature on the uptake of As(V) by ZnAlBi LDHs-embedded-Biene, b. Van’t Hoff plot for the uptake of As(V) by ZnAlBi LDHs-embedded-Biene

KD is referring to the uptake constant at equilibrium (L/g), R = general gas constant, T = temperature in (Kelvin), Ce (mg/L) is As(V) at equilibrium, qe is As(V) uptake by ZnAlBi LDHs-embedded-Biene (mg/g) (Dudek and Kołodyńska 2022). When lnKD is plotted versus 1/T as illustrated in Fig. 7b, (ΔS°) and (ΔH°) parameters can be identified (Table 7). Consequently, the negative (ΔH°) confirms that the uptake interaction of As(V) onto ZnAlBi LDHs-embedded-Biene was based on an exothermic reaction (Simsek et al. 2022). Moreover, the negative (ΔS°) indicates more controlled and regulated molecules at the liquid–solid interface during As(V) uptake operation (Islam et al. 2023), and the negative ΔG° refers to spontaneous uptake reaction in addition to the elevation in ΔG° values with enhancing temperature lead to decline in the uptake capacity (Simsek et al. 2022).

Table 7 Thermodynamic parameters for As(V) uptake by ZnAlBi LDHs-embedded-Biene at various temperatures

Application of ZnAlBi LDHs-embedded-Biene in As(V) uptake from actual water samples.

Several real water specimens tap water, sea water and wastewater were examined to confirm the efficiency of ZnAlBi LDHs-embedded-Biene for As(V) uptake from real water samples. It was found that the uptake percentages of As(V) from tap water were 93.29% and 90.52% by using 5 and 10 mg L−1As(V) for the first run, while the uptake percentage 90.17% was established at 15 mg L−1 As(V) as the result of the second run. Further, the As(V) uptake from sea water was 91.34% (5 mg L−1 As(V)), which is based on the first run, while 90.99% and 90.11% were accomplished by 10 and 15 mg L−1 As(V) according to the second run. Finally, the uptake percentages of As(V) from wastewater were 93.55% and 91.05% using 5 and 10 mg L−1 As(V) at the first run, while the uptake percentage was 90.09% for 15 mg L−1 as the result of second run. Thus, these results prove the excellent uptake and efficient performance of ZnAlBi LDHs-embedded-Biene for As(V) ions were captured from various water matrices.

Regeneration and recyclability of ZnAlBi LDHs-embedded-Biene

The investigated ZnAlBi LDHs-embedded-Biene was reutilized and regenerated to uptake As(V) for several repeated cycles to detect its efficiency and applicability. A 0.1 HCl solution was applied for the As(V) desorption from the nanosorbent surface for five cycles. The uptake results by using 5 mg L−1 As(V) were after first cycle 91.47%, second cycle 89.09%, third cycle 85.62%, fourth cycle 80.251% and the fifth cycle produced 74.09%. So, the recycling results indicate the stability of nanosorbent, applicability and excellent efficiency after five regeneration cycles (Ren et al. 2020).

Conclusion

ZnAlBi LDHs-embedded-Biene was simply assembled with a variety of exterior accessible functional groups including metal oxyhydroxide, metal oxides along with internal anions and mesoporous crystalline structure to favor uptake and elimination of As(V) with superior efficiently. As(V) was optimally removed from aqueous solutions at pH 3, in 30 min and 10 mg nanosorbent with 94.67%, 90.74% and 84.82% uptake percentages by 5, 10 and 25 mg L−1 As(V). Also, the ionic strength proved that the uptake percentages were not strongly affected by coexisting salts indicating a higher selectivity of ZnAlBi LDHs-embedded-Biene based on both electrostatic attraction mechanism with the exterior functional groups and ion-exchange between internal anion with As(V) ions. The thermodynamic evaluation detected that the interaction between As(V) and nanosorbent was exothermic and spontaneous in nature. The kinetic pattern indicated best-fitting expression via pseudo-second order to describe the As(V) interaction with ZnAlBi LDHs-embedded-Biene. Moreover, the evaluated adsorption isotherms indicated that the nanosorbent surface contributed and participated in As(V) uptake mechanism via heterogeneous multi-laminates and accessible functional groups. ZnAlBi LDHs-embedded-Biene exhibited excellent stability toward regeneration by HCl solution providing 91.47% As(V) adsorptive uptake result after the first cycle. ZnAlBi LDHs-embedded-Biene provided excellent adsorptive uptake process of As(V) ions as 93.29% (5 mg L−1) and 90.52% (10 mg L−1) As(V) from tap water were, while 93.55% (5 mg L−1) and 91.05% (10 mg L−1) As(V) wastewater. Thus, these results prove the excellent uptake and efficient performance of ZnAlBi LDHs-embedded-Biene for As(V) ions capturing from various water matrices.