Electrochemical Synthesis of SnO2/Al2O3 Nanocomposite: Characterization and Antibacterial Activity Study

Electrochemical Synthesis of SnO2/Al2O3 Nanocomposite: Characterization and Antibacterial Activity Study

Hayder Khudhair Khattar* Athmar Ali Kadhim Emad Abbas Jaffar Al-Mulla 

College of Health and Medical Techniques, Al-Furat Al-Awsat Technical University, An-Najaf 54001, Iraq

College of Medical Technology, The Islamic University, An-Najaf 54001, Iraq

Corresponding Author Email: 
haider.khattarckm11@atu.edu.iq
Page: 
531-540
|
DOI: 
https://doi.org/10.18280/acsm.490509
Received: 
16 September 2025
|
Revised: 
17 October 2025
|
Accepted: 
25 October 2025
|
Available online: 
31 October 2025
| Citation

© 2025 The authors. This article is published by IIETA and is licensed under the CC BY 4.0 license (http://creativecommons.org/licenses/by/4.0/).

OPEN ACCESS

Abstract: 

A simple electrochemical method was used to synthesize an SnO2/Al2O3 nanocomposite with a 1:1 ratio, by varying the size of the anode and examining the product using EDX analysis. This nanocomposite was then subsequently calcined at 200℃ to remove impurities and achieve a homogeneous crystal structure. The SnO₂/Al₂O₃ nanocomposite was synthesized using a surfactant (PVA) as a stabilizer in the electrodeposition method and characterized using techniques. X-ray diffraction (XRD) confirmed the crystalline phases with an average crystallite size of 28 nm, while field emission scanning electron microscopy (FESEM) revealed uniform nanoparticles with an average size 21-37 nm. High-resolution transmission electron microscopy (HRTEM) average size 20 nm, (UV-Vis) spectroscopy indicated a direct band gap in the range of 3.6 eV. The anti-bacterial activity of SnO₂/Al₂O₃ nanocomposite was also evaluated against Gram-positive Staphylococcus aureus and Streptococcus mutans, as well as Gram-negative Escherichia coli and Proteus mirabilis. the inhibition zone diameters increased with concentration from 10-1000 ppm, reaching 23 nm for S. aureus and 24 nm for Str. mutans at 1000 ppm, while E. coli and P. mirabilis reached 17 nm and 16 nm, respectively. these results show that SnO2/Al2O3 nanocomposite possesses structural, optical, and antibacterial properties, making it a suitable candidate for applications in electronics, optoelectronics, and antimicrobial coating.

Keywords: 

electrochemical, SnO2/Al2O3 nanocomposite 1:1 ratio, antibacterial activity

1. Introduction

behavior from the bulk states such as Al2O3, and SnO2. Because of the importance of these materials in multiple applications such as sensors, optoelectronic devices, solar cells, and thin-film transistors, their properties have been extensively studied [1, 2]. Nanometal oxides and their composites are well-known for their structural and high optical transmittance in the UV-Vis region and high IR reflectance [3-6]. The structural and visual properties are improved by relying on composition and size, as well as on the superficial area of nanoparticles. Therefore, the composite of hard nanostructured materials, such as SnO2 and Al2O3. It may be promising to produce excellent nanocomposites, there are many methods for synthesizing NPs, such as chemical vapor deposition [7], solvothermal methods [8], and microemulsion methods [9]. Sol-gel: this wet chemical technique involves converting precursors into a gel, which is then dried and sintered to form the desired nanomaterial. While relatively simple, it can require long processing times and the use of surfactants or stabilizers [10]. Laser ablation: this technique uses a laser to vaporize a target material, creating nanoparticles in a liquid medium. It can produce high-quality nanoparticles, however, the equipment can be expensive, and scaling up production can be challenging [11]. Hydrothermal methods utilizing high-pressure and high-temperature water for chemical reactions can be effective for synthesizing SnO2/Al2O3 materials, often require complex setups and specific conditions. The process depends on the solubility of reactants in water under elevated temperatures and pressures within a closed vessel. Among the most prominent limitations is the difficulty of ensuring a homogeneous nanodistribution of the two different SnO₂-Al₂O₃ phases within the compound, which leads to aggregation of particles or a decrease in the effectiveness of the reaction; the need for high temperatures or pressures or the use of expensive chemicals or harmful solvents; the slow process of synthesis or the inability to effectively expand industrially; weakness in controlling the fine nanostructure and controlling the size of particles and pores, which reduces the antibacterial or photocatalytic effectiveness [12]. The use of electrochemical deposition technology has emerged as a promising option to overcome these limitations due to the better control it provides over the distribution of the nanophase and the formation of a homogeneous film layer on the substrate or in the liquid; the possibility of performing synthesis at lower temperatures and with relatively simple equipment, which reduces the cost and facilitates expansion; greater speed in the formation and growth of particles with the possibility of adjusting the current density, voltage, time, and solvent to obtain a controlled particle size; and the opportunity To obtain more contact interfaces between the SnO₂-Al₂O₃, which may improve the antibacterial properties through greater surface activation or improved charge transfer. Through the data, the need has been to find a way through which effective and low-cost, as well as developmental and environmental, electrodeposition is a process that enables the formation of solid deposits on the surface of the electrically connected materials. It involves coating the surface of an electrode with a thin layer of the desired material. Electroplating employs an electric current to reduce dissolved metal cations in solution, forming a thin coating on the surface of the electrode. These deposits are produced using electrochemical cells, which consist of two conductors or semiconductors [13]. The electrochemical deposition mechanism involves immersing electrodes of the desired metal in a solution containing strong electrolyte ions, then applying an electric voltage or current that causes the ions to move toward the corresponding electrodes, where they undergo reduction and oxidation reactions. The anode electrode undergoing electrolysis is deposited in the middle of the solution by adjusting parameters such as voltage or current, deposition time, and ion concentration, where the size and shape of the particles can be controlled. However, this technique is rarely used in the preparation of SnO₂/Al₂O₃ nanocomposites. And evaluate its antibacterial effectiveness, which constitutes a research gap that this study seeks to fill by preparing the compound by electrochemical deposition, characterizing it, and evaluating its antibacterial activity. This study focuses on the synthesis of SnO₂/Al₂O₃ nanocomposite using an effective and surfactant-free electrodeposition method. The research aims to take advantage of the priority of electrical expression, such as the exact control of the thickness of the film, uniformity, and formation to produce these nanoparticles. The study will investigate the structural and optical properties of the synthesized nanocomposite, with the aim of exploring their potential applications in electronics, optoelectronics and sensor technologies, and antibacterial properties. The advantages of electrodeposition include its low cost, low energy consumption, environmentally friendly nature, reduced material waste, and the ability to produce thin films at low temperatures. The film formation process and fundamental properties can be controlled by adjusting the applied potential [14]. This study aims to prepare SnO₂/Al₂O₃ nanocomposite by electrophoretic deposition technique, describe its compositional, morphological, and surface properties, and then evaluate its antibacterial effectiveness. To achieve this by preparing nanoparticles or SnO₂-Al₂O₃ layers by controlling electrochemical deposition parameters such as voltage, current, time, and ion concentration, and analyzing the crystalline structure, shape, and phase distribution of the compound using techniques such as UV, FTIR, XRD, FESEM, EDX, and HRTEM; evaluating the antibacterial activity of the prepared compound against selected strains; and studying the effect of the concentration of the prepared nanocomposite.

2. Experiment

2.1 Materials and methods

The electrical system for preparing Nanocomposites SnO2/Al2O3 using a glass cell with a capacity of 100 ml and using the positive electrode, (anode), which is an Aluminum plate with an area of 2 cm2 (99%, Aldrich), as well as the Tin electrode 99% purity, with an area of 2cm (99%, Aldrich), the cathode negative electrode, which is graphite with an area of 4 cm, and the electrical current preparation The device used in the operation, is (5000 milli-ampere mA) maximum currents and 30 volts (V) maximum voltage, China) used in this method as a magnetic stirrer to mix the materials by stirring. Wash the poles used in this method with ethanol, then with deionized water, to remove the stranded organic materials. Add about 1 gram of a strong electrolyte material like KCL (99% Thomas, Indi). And 10 mL of 1g/100 mL of the stabilizer (Polyvinyl alcohol (PVA CDH-India), and deionized water, net foil of Al-Sn (4 cm × 2 cm × 0.25 mm) for every, inert graphite electrode (4 cm × 2 cm) was utilized as anode and cathode respectively, as shown in Table 1, They are positioned facing each other in the electric cell solution with a depth of 2 cm for both. The electric cell was operating for one hour for 40 to 50℃, It was centrifuged through a centrifugal device, The sediments were washed with deionized water and ethanol three times and dried at a temperature of 60℃ for one hour and after that, the sediments have then calcined in an oven at 200℃ for 2 hours, SnO2/Al2O3 composites were characterized by EDS, XRD, FESEM, HRTEM.

Table 1. The experimental data account for the EDS measurements

 

Chemicals

EDX

 

0.1 g/100 ml

KCL (99% Thomas, Indi)

 

0.5 g/100 ml

KOH (99% Thomas, Indi)

1g/100 mL. We take 10 ml

PVA (CDH-India)

0.7 μS/cm

Deionizedwater

10-12

pH

40-50℃

Temperature

500 mA/m2

Current density

1

24 cm2

Al (2 cm2)/Sn (2 cm2) Anode area

Al = 27.24

Sn = 31.89 (Atomic % ratio = 1:1.17)

O = 40.87

24 cm2

Graphite cathode

0.75-1.2 A

Current density

7-15 V/60 (min)

Voltage/time

2

18 cm2

Al (2 cm2)/Sn (1 cm2) Anode area

Al = 22.47

Sn = 43.65 (Atomic % ratio = 1:1.44)

O = 47.01

24 cm2

Graphite cathode

0.75-1.2 A

Current density

11-15 V/60 (min)

Voltage/time

3

15 cm2

Al (2 cm2)/Sn (0.5 cm2) Anode area

Al = 31.34

Sn = 21.84 (Atomic % ratio = 1:1.95)

O = 33.70

24 cm2

Graphite cathode

0.36-1.2 A

Current density

10-15 V/60 (min)

Voltage/time

4

18 cm2

Al (1 cm2)/Sn (2 cm2) Anode area

Al = 30.68

Sn = 23.31 (Atomic % ratio = 1:1.31)

O = 46.01

24 cm2

Graphite cathode

0.35 -1.2 A

Current density

7-15 V/60 (min)

Voltage/time

24 cm2

Al (2 cm2)/Sn (2 cm2) Anode area

5

15 cm2

Al (0.5 cm2)/Sn (2 cm2) Anode area

Al = 28.67

Sn = 28.32 (Atomic % ratio = 1:1.01)

O = 43.01

24 cm2

Graphite cathode

0.75 -1.2 A

Current density

7-15 V/60 (min)

Voltage/time

2.2 Synthesis of SnO2/Al2O3 nanocomposite (1:1 ratio)

The basis for the procedure is to control the size of SnO₂ and Al₂O₃ nanoparticles by adjusting the size of the electrodes used, especially the positive electrode anode, Al (2 cm²) and Sn (1 cm²), with an total anode area of 24 cm2 two faces, opposite graphite cathode 24 cm2, in the first experiment, we used two poles of equal size, and the result, according to Energy Dispersive X-Ray Spectroscopy (EDX) analysis (BRUKER) of the granules, was 1:1.17 atomic%. KCL, KOH, deionized water, and the PVA stabilizer were also used as the electrolyte. In the second experiment, the Sn and the Al poles were prepared with the required dimensions before the process, Al (2 cm²) and Sn (1 cm²), giving an anode area of 18 cm², opposite a graphite cathode 24 cm². The grain size ratio from EDX was 1:1.44 atomic%, as shown in Table 1. The change in this experiment involved only the anode size, while all other parameters remained fixed. In the third experiment, the tin pole area was 0.5 cm2 and the aluminum electrode area was 2 cm2, resulting in an EDX atomic% ratio for SnO2/Al2O3 of 1:1.95. In the fourth experiment, the tin pole size was 2 cm2 and the aluminum electrode size was 1 cm2, resulting in an EDX atomic% ratio of SnO2/Al2O3 was 1:1.31. The best result was obtained in the fifth experiment, using a tin pole area of 2 cm2 and an aluminum electrode size 0.5cm2, yielding an EDX atomic% ratio for SnO2/Al2O3 of 1:1.01. This ratio is considered the most optimal among the tested configurations and could be applied in various industrial and biological fields.

3. Results and Discussion

3.1 X-Ray Diffraction

The XRD patterns of the SnO2/Al2O3 nanocomposites are shown in Figure 1. The miller indices (hkl) of the diffraction peaks of SnO2/Al2O3 and nanocomposites were matched with JCPDS cards No. 71-1123 and No. 41-1445. The crystallite size was calculated using the Scherer formula [15, 16]. where D is the average crystallite size, λ is the used X-ray wavelength and k = 0.94 is a constant, θ is the diffraction angle (in degrees), and β is the full width at half maximum (FWHM, in radians) of the observed diffraction peak [17]. The XRD pattern of SnO2/Al2O3 in Figure 1 shows low peaks at 2θ values of 25.12°, 43.80°, and 57.60°, which correspond Al2O3 phase and confirm the presence of Al2O3 in the composite. Three intense diffraction peaks at 27.38°, 39.57°, and 49.24° can be attributed to the (110), (200), and (211) planes of SnO2, respectively. Compared with pure SnO2, the peaks near 27.38° in the SnO2/Al2O3 composites appear broader, which may be due to the overlap of the SnO2 (110) and Al2O3 (012) peaks [18, 19]. The tabulated X-ray data include: 2-Theta (diffraction angle), d (nm) (interplanar spacing), BG (background), height (peak intensity), I% (relative intensity), area (integrated peak intensity), I%2 (another relative intensity), full width at half maximum (FWHM), and crystal size (nm). The crystal size, calculated using the Scherrer equation, was approximately 11.13 nm as shown in Table 2.

Figure 1. XRD Pattern of SnO2/Al2O3 nanocomposites (a) Miller Indices (hkl), (b) Two-theta (2θ) represent angles related to the diffraction of X-rays by a crystalline sample

Table 2. The data obtained from X-ray diffraction (XRD)

2-Theta

D (nm)

BG

Height

I%

Area

I%2

FWHM

Crystal Size (nm)

10.802

0.81839

7

124

11.9

857

42.3

1.175

7

14.393

0.61487

0

96

9.2

776

38.3

1.374

6

18.793

0.47179

0

86

8.3

547

27

1.081

8

21.028

0.42213

0

124

11.9

833

41.1

1.142

7

25.12

0.35421

23

103

9.9

701

34.6

1.157

7

27.384

0.32542

60

1041

100

2027

100

0.331

28

33.203

0.2696

18

115

11

919

45.3

1.359

6

35.998

0.24928

0

95

9.1

489

24.1

0.875

10

39.573

0.22755

0

592

56.9

1528

75.4

0.439

21

43.822

0.20642

0

100

9.6

543

26.8

0.923

10

49.241

0.18489

0

277

26.6

906

44.7

0.556

17

57.614

0.15985

0

111

10.7

669

33

1.025

9

61.007

0.15175

0

69

6.6

470

23.2

1.158

8

65.439

0.14251

0

157

15.1

652

32.2

0.706

14

72.808

0.12979

0

99

9.5

653

32.2

1.121

9

3.2 UV-Vis spectrum

The UV-Vis spectra were analyzed to investigate the optical properties. The compounds and their absorbance changes are shown in Figure 2(a). The SnO2/Al2O3 nanocomposites exhibit strong absorption at 200 nm, with the absorption decreasing slightly as the wavelength increases. In the UV-visible range, no distinct peak was observed in the long wavelength region. Moreover, the maximum absorption was detected at about 235 nm, after which the absorbance gradually decreased [20]. The UV−Vis diffuse reflectance spectra display the results for the SnO2/Al2O3 samples in the range of 200-700 nm. The results showed that the absorption edge of the SnO2/Al2O3 nanocomposite occurred at 310 nm, which is in good agreement with the value (298 nm) reported by Mishra et al. [5] for SnO2-Al2O3 composites prepared via the sol−gel method. From the reflection spectra, the bandgap energy (Eg), was determined using a Tauc plot. In this method, (αhν)1/2(eV. cm–1)1/2 is plotted against hν, where α is the absorption co-efficient, h is Planck's constant, ν is the photon frequency, and γ depends on the type of electronic transition (e.g., 1/2 for indirect transition, 2 for direct transition). Extrapolating the linear portion of the curve to the x-axis (energy axis) gives the band gap energy, as shown in the inset of Figure 2(b). The bandgap values are found to be 3.65 eV for the SnO2/Al2O3 nanocomposite [21].

Figure 2. (a) UV–Vis Spectra, (b) the plot of (αhν)1/2 vs. photon energy (hν) for SnO2/Al2O3 nanocomposites

Figure 3. FTIR spectrum of the SnO2/Al2O3 nanocomposites

3.3 FTIR analysis of 1:1 SnO2/Al2O3 nanocomposites

For important information about functional groups in SnO2/Al2O3, FTIR spectra are recorded in transmission mode. FTIR spectra are shown in Figure 3. The broad peak at around 3200 cm-1 can be attributed to the O-H. It is noted that the lowest is in SnO2/Al2O3. Moreover, the bending vibrations of the O-H stretching vibration. It is noted that this peak is less intense in SNO2/Al2O3. Moreover, the bending vibrations of the O-H group are responsible for the peaks at about 1630 cm-1 in SnO2/Al2O3 nanocomposites [22]. The peak at about 500 cm-1 in SnO2/Al2O3 nanocomposites can be assigned to Sn-O or O-Sn-O vibrations [23]. The intensity of this peak decreases significantly for SnO2/Al2O3 nanocomposites, as shown in Figure 3. This decrease in metal-oxygen vibrations can be attributed to Sn or O vacancies in SnO2/Al2O3 powder. Similarly, the peaks at about 650 cm-1, 1100 cm-1, and 1450 cm-1 in SnO2/Al2O3 nanocomposites are mostly due to Al-O and O-Al-O vibrations as in Table 3 [24]. The disappearance or reduction of Al or O atom can lead to a decrease in the strength of these peaks. The absence of Sn, Al or O may result in weaker metal-oxygen stretching in the SnO2/Al2O3 bonding. The broadening of these peaks in the SnO2/Al2O3 powder suggests the presence of multiple weak vibrations instead of one strong vibration [22, 23]. From the FTIR spectrum, it is evident that the SnO2/Al2O3 powder contains a significant number of Sn, Al or O vacancies. However, these vacancies are relatively few compared to what might be expected, likely due to the absence of severe metal-oxygen loss during high-temperature calcination. Since the powder was dried in a hot air oven at 200℃ in an oxygen-adequate environment, the possibility of O vacancies is minimized. Furthermore, in SnO2/Al2O3 powder, the reduction of peak intensities related to metal-oxygen vibrations supports the presence of such lattice defects.

Table 3. FTIR vibration bands and their corresponding bond assignments for the SnO2/Al2O3 nanocomposite

Wavenumber (cm⁻¹)

Bond / Functional Group

Comment

Reference

3500–3400

O–H stretching (surface hydroxyl groups, adsorbed moisture)

Broad peak due to hydrogen-bonded OH groups

[25, 26]

3240–3435

O–H stretching (alcohol/water)

Strong hydrogen bonding in surface hydroxyls

[25]

1630–1640

H–O–H bending vibration

Bending mode of adsorbed water molecules

[26]

605 and 480

Sn–O and Sn–O–Sn stretching,

Asymmetric and symmetric

[27]

3400, 2925, 2851 and 1636

SnO2 synthesized

Indicating the presence of extract molecules in the SnO2 synthesized

[27]

1190–1200

C=O stretching (Sn–OH related)

Surface hydroxyl-carbonyl-like vibrations

[28]

860–820

O–Al–O stretching

Al₂O₃ lattice vibrations

[29]

620–600

Sn–O and Sn=O stretching

Characteristic of SnO₂

[30]

560–500

O–Al–O bending / Sn–O–Sn symmetric stretching

Evidence of mixed oxide bonding

[29, 30]

3.4 FESEM, RHTEM, and EDX

Figure 4 shows the field emission scanning electron microscopy (FESEM) image of SnO2/Al2O3 nanocomposite particles in various size ranges up to 500 nm. The microscopic image reveals that the nanocomposites tend to aggregate on the surface, confirming the presence of pores, which can be useful in sensing applications. The particle size is approximately 21-37 nm, which is larger than that calculated from XRD analysis [31]. Figure 5 shows high-resolution transmission electron microscopy (HR-TEM) images of the SnO2/Al2O3 nanocomposite. HRTIM analysis is among the most effective methods to investigate nanoparticle size. The practical size in this work was found to be around 20 nm. The images also show that the grain size distribution is not uniform. During sample preparation, agglomeration of particles may occur. Energy dispersive X-ray (EDX) spectroscopy of the SnO2/Al2O3 nanocomposite confirms the presence of its constituent elements. Energy‑Dispersive For the 1:1 sample (Experiment 5), the mass composition was O = 43.01%, Al = 28.67% and Sn = 28.32%. Also, maps with a resolution of 768 × 512 were obtained to ensure racial distribution. These results confirm the expected chemical structure of the SnO₂/Al₂O₃ compound, and show good agreement with X-Ray Diffraction (XRD) data, as in Figure 6 [32, 33].

Figure 4. FESEM micrograph of SnO2/Al2O3 nanocomposite

Figure 5. HRTEM micrograph of SnO2/Al2O3 nanocomposite

The average particle size of SnO₂/Al₂O₃ nanoparticles, as determined by XRD, FE-SEM, and HRTEM analyses, showed a good agreement through the three characterization techniques, with minor differences attributed to the related differences in the principles of measurement. The XRD-derived size, calculated using the Scherrer equation, reflects the average size of crystals, and that may be a little smaller than the size of the particles observed in the FESEM and HRTEM images due to the potential assembly and the polycrystal nature of the particle. FESEM analysis presented a direct visualization of particle morphology. This reveals a structure with agglomeration and uniform particles with a small mass. HRTEM has provided high-resolution images, allowing an accurate measurement of individual molecules and confirming the crystallinity of the lattice corresponding to SnO₂ and Al₂O₃, when compared to the values mentioned in the literature shown in the Table 4, the sizes of the particles obtained within the expected range of XRD were within the expected range of 11.13 nm, in line with pre-reported studies [5, 34-37]. Slight deviations from literature data may result from differences in the terms of synthesis, base materials, calcination temperatures, and SnO₂ to Al₂O₃ ratios. This indicates that the approved synthesis method produces effectively from nanoparticles of similar particles, or smaller than those that have been previously reported, which can enhance the surface area and catalytic activity.

Figure 6. Energy dispersive X-ray (EDX) spectroscopic analysis of SnO2/Al2O3 nanocomposite showing Al=28.67 atomic%, Sn =28.32 atomic%, O = 43.01 atomic%

Table 4. Comparing the average particle size of the SnO₂/Al₂O₃ nanoparticles calculated by XRD, FESEM, and HR-TEM techniques with reported literature values

Size Technique

Size (nm)

Shape/Surface

Reported Values in Literature (nm)

Literature Reference

XRD

11.13

Derived from Scherrer equation

13-15

12.1

[34]

[37]

[5]

FESEM

21-37

Morphology observation

62

[5]

HRTEM

20-60

Detailed shape and boundary analysis

20-25, 30

20-80

5-100

[35]

[36]

[37]

4. Antibacterial Activity Using Agar Rout

The in vitro antibacterial activity of the compound was evaluated against Staphylococcus Aureus, Streptococcus mutans, Escherichia coli, and Proteus mirabilis using the well-diffusion method on Muller-Hinton agar (7). Wells of 6 mm in diameter were made in the agar plates using a sterile cork borer. The agar surfaces were then inoculated with each bacterial strain. The synthesized SnO2/Al2O3 nanocomposite particles were dissolved in dimethylformamide (DMF) to prepare solutions with concentrations of 10, 100, and 1000 ppm. The plates were incubated at 37℃ for 12 hours, and the inhibition zones were measured in millimeters. Each test was performed in triplicate, and the average inhibition zone’s diameter was recorded.

4.1 Antibacterial activity studies

Semiconductors or metal oxides are among the promising alternatives for antibacterial application. after incubation, the wells loaded with SnO2/Al2O3 nanocomposite showed bactericidal activity against E. coli and P. aeruginosa, comparable to Ciprofloxacin (used as the standard), as evidenced by the inhibition zones. The antibacterial capacity is strongly influenced by the morphology of nanoparticles. Smaller nanoparticles are generally preferred few sizes, these nanoparticles are preferred because of their potent bactericidal effects. Several reports indicate that nanoparticles in contact with bacterial walls can inhibit the synthesis of sulfur or phosphorus-containing envelope proteins. When the DNA interacts with metal oxide ions, the Al+3 and Sn+2 ions may induce biochemical changes that ultimately lead to cell death. This may involve interference with the electron transport chain (ETC), disruption of ion pumps alteration of ion movement across the membrane [38]. Weakening of the bacterial plasma membrane can further stimulate cell death. Overall, the tested microorganism was not resistant and did not display any biofilm formation in response to the synthesized nanoparticles. These findings are consistent with previous studies, supporting the bactericidal potential of Al2O3 and SnO2 nanoparticles.

Table 5. Antibacterial activity of SnO2/Al2O3 at various concentrations, expressed as the zone of inhibition mm

Concentration

SnO2/Al2O3

G(+Ve)

G(-Ve)

S. aureus

Str. mutans

E. coli

P. mirabilis

10 ppm

12

15

7

9

100 ppm

17

20

14

12

1000 ppm

23

24

17

16

Figure 7. Statistical representation of antibacterial activity of SnO2/ Al2O3

The biological activity of the (SnO2/Al2O3) nanocomposite was evaluated in vitro against S. aureus, Str. mutans, E. coli, and mirabilis using the well-diffusion method. The inhibition zones for each bacterial strain are presented in Table 5, with statistical analysis is shown in Figure 7. The results indicate that the (SnO2/Al2O3) Gram-positive bacteria has a stronger effect on the Gram-positive bacteria (S. aureus, Str. mutans) than on the Gram-negative strains tested, under identical experimental conditions. At concentrations of 10 and 100 ppm, the nanocomposite exhibited mild activity against (S. aureus, Str. mutans), while at 1000 ppm it showed high antibacterial potency. In contrast, its activity against E. coli and P. mirabilis was moderate to weak at 10 and 100 ppm, and moderate at 1000 ppm. The inhibition effect at the concentration of 1000 ppm was moderate. These findings are consistent with previous reports [39], which also observed higher inhibition of Gram-positive bacteria compared to Gram-negative bacteria, despite the thicker peptidoglycan layer of Gram-positive. This difference is likely due to the outer membrane of Gram-negative bacteria, which acts as an additional barrier to nanoparticle penetration [40]. The results in Table 5 further support this, showing greater efficacy against the Gram-positive strains. This can be explained by structural differences in the bacterial cell envelope gram-negative bacteria possess an extra outer membrane in addition to the peptidoglycan layer, creating an additional barrier that limits biocide penetration [41].

4.2 The experimental determination of the minimum inhibitory concentration (MIC) for antibacterial activity

As shown in Table 6, these results provide evidence that the nanocomposite SnO2/Al2O3 has inhibitory effects against the studied bacterial strains, Staphylococcus aureus and Streptococcus mutans, were inhibited at a concentration of 10 ppm, whereas Escherichia coli and Proteus mirabilis were inhibited at a concentration of 100 ppm. Thus, the results confirm that Gram-negative bacteria are approximately ten times less sensitive than Gram-positive bacteria [42, 43].

Table 6. Shows the results of the minimum inhibitory Concentration (MIC) of the SnO2/Al2O3 nanocomposite

MIC (ppm)

Bacterial Strain

10

Staphlococcus aureus (S. aureus)

10

Streptococcus mutans (Str. mutans)

100

Escherichia coli (E. coli)

100

Proteus mirabilis (P. mirabilis)

4.3 Antibacterial activity inhibition mechanism by SnO₂/Al₂O₃ nanocomposite

The low sensitivity of Gram-negative bacteria is likely due, at least in part, to the limited absorption of nanoparticles and to differences in the structure of the cell wall, an outer layer of lipopolysaccharide (LPS) reduces the permeability of nanoparticles [44-46]. As for the antibacterial behavior of the SnO2/Al2O3 nanocomposite, several mechanisms are involved, including the generation of reactive oxygen species (ROS), which leads to the formation of a matrix that directly damages the cellular components of bacteria [37, 43, 44, 46]. Evidence indicates that the cell membrane was compromised, likely through direct interaction with the cell surface [43, 44], leading to the release of active ions that interfered with essential bacterial processes such as iron siderophore absorption [45, 46]. Therefore, the SnO2/Al2O3 nanocomposite could have practical applications in the medical and dental fields, such as antibacterial coatings for medical devices and oral care products [42, 47, 48]. It should be noted that Gram-negative bacteria exhibit relative resistance to this compound, indicating the need for targeted strategies to enhance its antibacterial efficacy.

5. Conclusion

Using the simple electrical method, the SnO2/Al2O3 nanocomposite was prepared. The crystal size and the morphology of the sample, analyzed via X-ray diffraction and several different techniques, indicate the products crystallinity, crystallite size, and other characteristic properties, signaling that the utilized process is proper and economical for obtaining SnO2/Al2O3 nanocomposites. It can be stated that SnO2/Al2O3 at a concentration of (1000) ppm can act as powerful antibiotics agent with potential effects on other bacteria types not yet tested. However, SnO2/Al2O3 may not show a similar effect in vivo despite the bacteria sensitivity observed in vitro, as in vivo conditions often differ significantly. Therapeutic process involves complex metabolic and physiological reactions that are difficult to estimate despite knowing the final treatment outcomes, whether positive or negative. Additionally, other factors may interfere with antibiotic efficacy, resulting in outcomes that are difficult to explain. It was observed that the Gram-positive bacteria are inhibited by relatively low concentration of the nanocomposite, whereas Gram-negative bacteria require much higher concentration for inhibition. The difference in sensitivity is due to variations in the general structure of the bacterial cell wall, with each bacterial type requiring a specific compound for effective inhibition. The nanocomposite has shown an additional advantage: antibacterial activity against antibiotic-resistant strains due to the multiple mechanisms, including increasing ROS generation, membrane disruption, and ion release. The nanocomposite can be used in many fields, such as antibacterial coatings and medical devices. It may also be applied in further studies, such as surface property modifications to improve its efficacy against Gram-negative bacteria. SnO₂/Al₂O₃ nanocomposite, a semiconductor with controlled capacity and porosity, has multiple enforcement possibilities and also has work in the industrial field. SnO₂ is supported by a stable Al₂O₃ core. This feature makes it suitable for fields such as gas sensors or catalytic materials. For example, a similar version has been studied for detecting low-concentration gases using small-sized oxides. At the environmental/biological level, preliminary studies have shown the antibacterial effectiveness of nano-metal oxides, including alumina, which opens up to use in sterilization or antimicrobial materials. Therefore, our study, which focuses on the synthesis of SnO₂/Al₂O₃ via electrochemical deposition, is not just an addition to knowledge but rather paves the way for potential applications in environmental and health sensors, that is, in a series linking nanopreparation and functional improvement.

  References

[1] Shah, N.A., Gul, M., Abbas, M., Amin, M. (2019). Synthesis of metal oxide semiconductor nanostructures for gas sensors. In Gas Sensors. IntechOpen. https://doi.org/10.5772/intechopen.86815

[2] Al-Abboodi, H., Fan, H., Al-Bahrani, M., Abdelhussien, A., Mohamad, B. (2024). Mechanical characteristics of nano-crystalline material in metallic glass formers. FACTA UNIVERSITATIS Series: Mechanical Engineering, 22(2): 315-328. https://doi.org/10.22190/FUME230128016A

[3] Abd-Elnaiem, A.M., Rashad, M., Hanafy, T.A., Shaalan, N.M. (2023). Improvement of optical properties of functionalized polyvinyl alcohol-zinc oxide hybrid nanocomposites for wide UV optoelectronic applications. Journal of Inorganic and Organometallic Polymers and Materials, 33(8): 2429-2444. https://doi.org/10.1007/s10904-023-02616-w

[4] Sohn, H.Y., Murali, A. (2021). Plasma synthesis of advanced metal oxide nanoparticles and their applications as transparent conducting oxide thin films. Molecules, 26(5): 1456. https://doi.org/10.3390/molecules26051456

[5] Mishra, N.K., Kumar, C., Kumar, A., Kumar, M., Chaudhary, P., Singh, R. (2015). Structural and optical properties of SnO2–Al2O3 nanocomposite synthesized via sol-gel route. Materials Science-Poland, 33(4): 714-718. https://doi.org/10.1515/msp-2015-0101

[6] Chavan, G.T., Kim, Y., Khokhar, M.Q., Hussain, S.Q., et al. (2023). A brief review of transparent conducting oxides (TCO): The influence of different deposition techniques on the efficiency of solar cells. Nanomaterials, 13(7): 1226. https://doi.org/10.3390/nano13071226

[7] Wu, F., Wang, C., Hu, H.Y., Pan, M., et al. (2019). One-step synthesis of hierarchical metal oxide nanosheet/carbon nanotube composites by chemical vapor deposition. Journal of Materials Science, 54(2): 1291-1303. https://doi.org/10.1007/s10853-018-2889-9

[8] Liang, Y., Jiang, L., Xu, S., Ju, W., et al. (2024). Synthesis and characterization of Fe3O4 nanoparticles prepared by solvothermal method. Journal of Materials Engineering and Performance, 33(13): 6804-6815. https://doi.org/10.1007/s11665-023-08431-1

[9] Asab, G., Zereffa, E.A., Abdo Seghne, T. (2020). Synthesis of Silica-Coated Fe3O4 nanoparticles by microemulsion method: Characterization and evaluation of antimicrobial activity. International Journal of Biomaterials, 2020(1): 4783612. https://doi.org/10.1155/2020/4783612

[10] Patel, G.H., Chaki, S.H., Kannaujiya, R.M., Parekh, Z.R., Hirpara, A.B., Khimani, A.J., Deshpande, M.P. (2021). Sol-gel synthesis and thermal characterization of SnO2 nanoparticles. Physica B: Condensed Matter, 613: 412987. https://doi.org/10.1016/j.physb.2021.412987

[11] Crivellaro, S., Guadagnini, A., Arboleda, D.M., Schinca, D., Amendola, V. (2019). A system for the synthesis of nanoparticles by laser ablation in liquid that is remotely controlled with PC or smartphone. Review of Scientific Instruments, 90(3): 033902. https://doi.org/10.1063/1.5083811

[12] Kafi, F.S.B., Gunaratna, B.H., Jayathilaka, K.M.D.C., Wijesundera, R.P. (2024). Electrodeposited thin film SnO2 photoelectrode for PEC applications. Proceedings of the Technical Sessions, 40(2024): 47-53.

[13] Pelegrini, S., Brandt, I.S., Cid, C.C.P., Isoppo, E.A., Viegas, A.D., Pasa, A.A. (2015). Electrochemical Cl doping of Cu2O: Structural and morphological properties. ECS Journal of Solid State Science and Technology, 4(7): P181. https://doi.org/10.1149/2.0041507jss

[14] Zahra, T., Alanazi, M.M., Alahmari, S.D., Abdelmohsen, S.A., et al. (2024). Hydrothermally synthesized ZnSe@ FeSe nanocomposite: A promising candidate for energy storage devices. International Journal of Hydrogen Energy, 59: 97-106. https://doi.org/10.1016/j.ijhydene.2024.01.293

[15] Kumar, S., Mote, V.D., Prakash, R., Kumar, V. (2016). X-ray analysis of α-Al2O3 particles by Williamson–Hall methods. Materials Focus, 5(6): 545-549. https://doi.org/10.1166/mat.2016.1345

[16] Nagarani, S., Sasikala, G., Satheesh, K., Yuvaraj, M., Jayavel, R. (2018). Synthesis and characterization of binary transition metal oxide/reduced graphene oxide nanocomposites and its enhanced electrochemical properties for supercapacitor applications. Journal of Materials Science: Materials in Electronics, 29(14): 11738-11748. https://doi.org/10.1007/s10854-018-9272-0

[17] Mahanta, R., Chetri, P., Shukla, N., Bora, D. (2024). Daylight photocatalytic activity of SnO2@ Al2O3 core-shell: An experimental & first principles study. Physica B: Condensed Matter, 690: 416230. https://doi.org/10.1016/j.physb.2024.416230

[18] Huang, X., Feng, Y., Bai, C., Wu, K., Ke, J., Xiong, D., He, M. (2020). SnO2-Al2O3-graphite nanosheets as a long-life and high-rate anode material for lithium-ion batteries. Chemical Physics Letters, 749: 137456. https://doi.org/10.1016/j.cplett.2020.137456

[19] Kumar, C., Mishra, N.K., Kumar, A., Bhatt, M., Chaudhary, P., Singh, R. (2016). Structural investigation of nanomixed xSnO2–Al2O3 synthesized by sol–gel route. Applied Nanoscience, 6(7): 1059-1064. https://doi.org/10.1007/s13204-015-0515-6

[20] Singh, R., Kumar, M., Shankar, S., Singh, R., Ghosh, A.K., Thakur, O.P., Das, B. (2015). Effects of Sb, Zn doping on structural, electrical and optical properties of SnO2 thin films. Materials Science in Semiconductor Processing, 31: 310-314. https://doi.org/10.1016/j.mssp.2014.12.010

[21] Sadeghzadeh-Attar, A. (2020). Binary Zn-doped SnO2/Al2O3 nanotube composites for visible-light-driven photocatalytic degradation of basic blue 41. ACS Applied Nano Materials, 3(10): 9931-9942. https://doi.org/10.1021/acsanm.0c01939

[22] Jabeen, A., Rahman, S.U., Shah, A., Khan, S.U., et al. (2025). Tailoring novel SnO2/α-MnO2 composites for photocatalytic performance under visible-light. Arabian Journal for Science and Engineering, 50(9): 6459-6470. ‏ http://doi.org/10.1007/s13369-024-09406-7

[23] Monga, K., Aggarwal, M., Singh, V., Chaudhary, S. (2024). Zinc and [Zinc, Nickel]: Co doped SnO2 nanoparticles as prospective electron transport layer materials for efficient lead free-MASnI3 perovskite solar cells. Chemical Physics Impact, 8: 100651.‏ https://doi.org/10.1016/j.chphi.2024.100651

[24] Dhawale, V.P., Khobragade, V., Kulkarni, S.D. (2018). Synthesis and characterization of aluminium oxide (Al2O3) nanoparticles and its application in azodye decolourisation. Chemistry, 27: 31. https://doi.org/10.11648/j.ijec.20180201.13

[25] Guo, X., Qing, Y., Wu, Y., Wu, Q. (2016). Molecular association of adsorbed water with lignocellulosic materials examined by micro-FTIR spectroscopy. International Journal of Biological Macromolecules, 83: 117-125. https://doi.org/10.1016/j.ijbiomac.2015.11.047

[26] Guo, X., Yuan, H., Xiao, T., Wu, Y. (2019). Application of micro-FTIR spectroscopy to study molecular association of adsorbed water with lignin. International Journal of Biological Macromolecules, 131: 1038-1043. https://doi.org/10.1016/j.ijbiomac.2019.03.193

[27] González, E.Q., Medina, E.L., Robles, R.V.Q., Gálvez, H.E.G., et al. (2022). A study of the optical and structural properties of SnO2 nanoparticles synthesized with Tilia cordata applied in methylene blue degradation. Symmetry, 14(11): 2231. https://doi.org/10.3390/sym14112231

[28] Giridasappa, A., Shareef, M.I., Gopinath, S.M., Rangappa, D., Shivaramu, P.D., Sabbanahalli, C. (2023). Synthesis, antioxidant, bactericidal and antihemolytic activity of Al2O3 and SnO2 nanoparticles. Proceedings of the National Academy of Sciences, India Section B: Biological Sciences, 93(4): 871-882. https://doi.org/10.1007/s40011-023-01444-9

[29] Pourmadadi, M., Hafezi, A.R., Mafirad, S., Nouri, M., et al. (2025). Development and evaluation of SA/PVP/γ-Al₂O₃ nanocarriers for targeted delivery of curcumin in colon cancer therapy. Journal of Drug Delivery Science and Technology, 105: 106632. https://doi.org/10.1016/j.jddst.2025.106632

[30] Tański, T., Smok, W., Matysiak, W. (2021). Characterization of morphology and optical properties of SnO₂ nanowires prepared by electrospinning. Bulletin of the Polish Academy of Sciences. Technical Sciences, 69(5): 1-8. http://doi.org/10.24425/bpasts.2021.137507

[31] Ghrib, B., Slama, S.S.B., Dimassi, W., Gaidi, M., Ouertani, R. (2021). Effect of Al2O3 coatings on microstructural and optoelectronic properties of porous Si/SnO2 composites. https://doi.org/10.21203/rs.3.rs-670960/v1

[32]  Caputo, F., Clogston, J., Calzolai, L., Rösslein, M., Prina-Mello, A. (2019). Measuring particle size distribution of nanoparticle enabled medicinal products, the joint view of EUNCL and NCI-NCL. A step by step approach combining orthogonal measurements with increasing complexity. Journal of Controlled Release, 299: 31-43. https://doi.org/10.1016/j.jconrel.2019.02.030

[33] Jayachandran, V., Dhandapani, V.S., Muniappan, E., Park, D., Kim, B., Arun, A.P., Ayyappan, P.R. (2022). Assessment of the synergetic performance of nanostructured CeO2-SnO2/Al2O3 mixed oxides on automobile exhaust control. Materials, 15(23): 8460. https://doi.org/10.3390/ma15238460

[34] Oliveira, M.L., Da Boit, K., Schneider, I.L., Teixeira, E.C., Borrero, T.J. C., Silva, L.F. (2018). Study of coal cleaning rejects by FIB and sample preparation for HR-TEM: Mineral surface chemistry and nanoparticle-aggregation control for health studies. Journal of Cleaner Production, 188: 662-669. https://doi.org/10.1016/j.jclepro.2018.04.050

[35]  Zhu, S., Liu, J., Sun, J. (2019). Precise growth of Al2O3/SnO2/CNTs composites by a two-step atomic layer deposition and their application as an improved anode for lithium ion batteries. Electrochimica Acta, 319: 490-498. https://doi.org/10.1016/j.electacta.2019.07.027

[36] Hussein, S.A., Gehad, S.M., Ebnalwaled, A.A. (2023). Effect of metal oxides ratio on the properties of SnO2/Al2O3 nanocomposites. IOSR Journal of Applied Physics, 14(1): 52-59. https://doi.org/10.9790/4861-1404015259

[37] Kim, Y.E., Lee, W., Youn, M.H., Jeong, S.K., Kim, H.J., Park, J.C., Park, K.T. (2019). Leaching-resistant SnO2/γ-Al2O3 nanocatalyst for stable electrochemical CO2 reduction into formate. Journal of Industrial and Engineering Chemistry, 78: 73-78. https://doi.org/10.1016/j.jiec.2019.05.042

[38] Rodriguez, R., Müller, S., Colombeau, L., Solier, S., Sindikubwabo, F., Cañeque, T. (2025). Metal ion signaling in biomedicine. Chemical reviews, 125(2), 660-744.‏ https://doi.org/10.1021/acs.chemrev.4c00577

[39] Singer, L., Karacic, S., Szekat, C., Bierbaum, G., Bourauel, C. (2023). Biological properties of experimental dental alginate modified for self-disinfection using green nanotechnology. Clinical Oral Investigations, 27(11): 6677-6688. https://doi.org/10.1007/s00784-023-05277-8

[40] Mukhanu, C. (2023). Screening of natural compounds inhibiting biofilm formation by antibiotic resistant enterobacteria using experimental trials and computational molecular modelling (Master's thesis, University of Pretoria (South Africa)).

[41] Díez-Pascual, A.M., Díez-Vicente, A.L. (2016). PEGylated boron nitride nanotube-reinforced poly (propylene fumarate) nanocomposite biomaterials. RSC Advances, 6(83): 79507-79519. https://doi.org/10.1039/C6RA09884C

[42] Garg, D., Chatterjee, T. (2023). A case of profound hypertriglyceridemia causing pseudohypobicarbonatemia. Cureus, 15(4): e37489. https://doi.org/10.7759/cureus.37489

[43] Yao, Z., Wu, D., Liu, J., Wu, W., Zhao, H., Tang, J. (2017). Recycling of typical difficult-to-treat e-waste: Synthesize zeolites from waste cathode-ray-tube funnel glass. Journal of Hazardous Materials, 324: 673-680. https://doi.org/10.1016/j.jhazmat.2016.11.041

[44] Mammari, N., Lamouroux, E., Boudier, A., Duval, R.E. (2022). Current knowledge on the oxidative-stress-mediated antimicrobial properties of metal-based nanoparticles. Microorganisms, 10(2): 437. https://doi.org/10.3390/microorganisms10020437

[45] Zhang, J., Chen, Z., Chen, Q. (2024). Advanced nano-drug delivery systems in the treatment of ischemic stroke. Molecules, 29(8): 1848. https://doi.org/10.3390/molecules29081848

[46] Franco, D., Calabrese, G., Guglielmino, S.P.P., Conoci, S. (2022). Metal-based nanoparticles: Antibacterial mechanisms and biomedical application. Microorganisms, 10(9): 1778. https://doi.org/10.3390/microorganisms10091778

[47] Swartjes, J.J., Sharma, P.K., Kooten, T.V., van der Mei, H.C., Mahmoudi, M., Busscher, H.J., Rochford, E.T. (2015). Current developments in antimicrobial surface coatings for biomedical applications. Current Medicinal Chemistry, 22(18): 2116-2129. https://doi.org/10.2174/0929867321666140916121355

[48] Brandelli, A. (2024). Nanocomposites and their application in antimicrobial packaging. Frontiers in Chemistry, 12: 1356304. https://doi.org/10.3389/fchem.2024.1356304