Mehrdad Zolfalizadeha,
Hadi Khalilnezhadb,
Saeid Nickabadib and
Behrang Golmohammadi*c
aFaculty of Chemical and Petroleum Engineering, University of Tabriz, Tabriz, Iran
bDepartment of Mechanic Engineering, Imam Khomeini University of Marine Science & Technology, Nowshahr, Iran
cDepartment of Physical Chemistry, University of Tabriz, Tabriz, Iran. E-mail: behrangrose@gmail.com
First published on 31st July 2025
Magnetic nanocomposite (MNC) colloids can be employed in guided systems under a magnetic field, as demonstrated in this work for the removal of Escherichia coli (E. coli) from water. Herein, we employed a magnetic reactor system incorporating an Fe3O4@ZIF-8@Ag2O nanocomposite colloid with destructive power against bacteria, focusing on elucidating the influence of key parameters, including the applied magnetic field and the presence of the MNCs, on the efficiency of bacterial removal. To assess bacterial viability, experiments were conducted in both phosphate-buffered saline (PBS) and a standard aqueous solution. The experimental design aimed to optimize bacterial removal by employing a 2 mT magnetic field, generated by a solenoid coil-assisted tubular reactor, which contributed to the stabilization of the magnetic nanoparticle (MNP) colloids. As the bacterial suspension traversed the reactor, interactions between E. coli and the nanoparticles resulted in collisions that induced cellular disruption and subsequent cell death. Under magnetic field conditions, the removal efficiency of E. coli increased by 29.6% after 60 minutes relative to control experiments without a magnetic field, culminating in an overall removal rate of 99.83% within the same period. Additionally, kinetic modeling using the Weibull function in the absence of a magnetic field yielded a minimal root mean square error (RMSE) of 7%, while the maximum standard deviation in the average E. coli concentration reached 75% at 40 minutes under magnetic conditions. These findings suggest that using stable MNC colloids under a controlled magnetic field markedly improves bacterial removal from water.
Magnetic colloidal nanocomposite systems have recently emerged as a promising solution for antibacterial applications.9–11 In this respect, Lin et al. demonstrated complete arsenic adsorption within 30 minutes using magnetic γ-Fe2O3 nanoparticles, with the adsorption process remaining stable across a pH range of 3 to 11 and unaffected by common anions.12 Han et al. reported that manganese oxide-coated zeolite (MOCZ) effectively removed copper(II) and lead(II) ions from aqueous solutions, exhibiting a stronger affinity for lead.13 Furthermore, Gonzalez-Vazquez et al. found that the application of an external magnetic field enhanced the removal rates of cadmium and zinc from activated carbon adsorbents.14 In microbial decontamination, water treatment plants have adopted methods such as membrane filtration, powdered activated carbon adsorption, sand filtration, and oxidation to eliminate E. coli.15–18 Complementary research has shown that advanced materials, including AgNO3-activated carbon composites and MNCs, offer substantial antibacterial activity, achieving removal efficiencies as high as 99.99% and enabling rapid disinfection at optimal concentrations.19–22
Auxiliary devices such as solenoid wire coils, which generate magnetic fields upon the passage of an electric current, have been employed to further enhance nanoparticle performance in water treatment systems.23 Magnetic fields not only influence the migration and distribution of nanoparticles, as demonstrated in studies on water/alumina nanofluids,24 but also improve the removal efficiency of contaminants in wastewater treatment processes, including the elimination of nitrogenous compounds and E. coli.25–27 Collectively, these studies illustrate an integrated approach that combines advanced materials with magnetic field applications to achieve more effective and rapid water purification, addressing both chemical and microbiological contaminants while supporting environmental and public health objectives.
In this study, an advanced column reactor system filled with MNC colloid has been developed for water disinfection, and was designed to improve the bacterial removal efficiency from aqueous media. This system incorporates Fe3O4@ZIF-8@Ag2O MNCs, which have demonstrated high removal efficiency even at low concentrations in a colloidal mixture of aqueous feed. There are, however, a few risks from the release of silver nanoparticles (AgNPs) into the environment. In essence, the increasing use of AgNPs across various consumer, industrial, and even agricultural sectors presents a tangible risk of environmental pollution. These nanoparticles can enter the environment through diverse pathways, including wastewater discharge, product disposal, and atmospheric deposition. Once in the environment, AgNPs pose a threat to a wide range of organisms, from aquatic life and soil microorganisms to potentially higher trophic levels through bioaccumulation. Their toxicity and the possibility of transformation into different, potentially harmful forms raise concerns about ecosystem disruption and long-term ecological consequences. Furthermore, the widespread release of these antibacterial agents could contribute to the growing problem of antibiotic resistance. While the full extent of their environmental impact is still under investigation, the presence of AgNPs in various environmental samples serves as a clear indicator of existing pollution. Therefore, continued research into their fate and effects, along with the development of strategies to minimize their release, is crucial to mitigate the potential environmental risks associated with the increasing use of AgNPs.28–35
The system was engineered to stabilize MNPs within the reactor, facilitating increased Brownian motion and enhancing interactions between nanoparticles and bacteria. We hypothesized that a magnetic field generated by a copper solenoid coil and supplied with direct current (DC) would stabilize the nanoparticles within the system and direct their movement along the coil's axis to isolate the effects of MNPs. The solenoid coil was specifically designed to minimize heat generation when a magnetic field was applied, which could reduce the E. coli removal. One advantage of this study over previous research in this field is the use of a novel antimicrobial nanostructure, guided by a new magnetic system, to enhance microbial removal efficiency.
A solenoid coil was uniformly wrapped around the column at equidistant intervals. Upon the application of a DC through the coil, a magnetic field was generated.
During the trial phase, the prepared contaminant water was pumped through the reactor at a flow rate of 10 mL per minute using a peristaltic pump. This flow rate was selected based on the reactor dimensions and the desired residence time of the bacteria within the reactor, thereby optimizing the contact between the bacteria and the nanoparticles to improve the bacterial removal efficiency. While the solution flowed downward through the reactor, a magnetic field was applied using a solenoid coil powered by a 12 V DC power supply at a current of 0.2 A. The treated water, referred to as the effluent, was collected from the top outlet of the reactor at specific time intervals (10, 20, 30, 40, 50, and 60 minutes) for sampling.
To determine the kinetic reaction parameters, the experimental results were fitted to the Chick–Watson and Weibull models. The Chick–Watson kinetic model is represented by the following equation:
log(Ce) = log(Ci) − kmaxt | (1) |
![]() | (2) |
The RMSE equation was given by Mecha et al., 2020, as follows:46
![]() | (3) |
![]() | (4) |
Solenoid wire coils were employed to generate a controlled magnetic field, a configuration that has found numerous applications in industry. In this study, the magnetic field generated by the coils surrounding the MNPs within the reactor was utilized to enhance particle stability in the bed, increase Brownian motion and particle collisions, and ultimately improve the bacterial removal efficiency. The system's effectiveness was further optimized by adjusting its parameters. The magnetic field was oriented to induce a counter-current flow of the Fe3O4@ZIF-8@Ag2O nanocomposite relative to the influent contaminant water. An increase in the magnetic field correspondingly reduced the velocity profile of the particles, which extended the contact time between the E. coli bacteria and the adsorbents, resulting in a higher percentage of E. coli removal in the effluent. Conversely, the Brownian motion of the particles within the reactor led to an increasing temperature profile and a decreasing concentration profile. Although the decreased concentration extended the reaction time, the elevated temperature posed challenges by potentially disrupting data measurement processes. Consequently, temperature control was deemed crucial, and the temperature profile was regulated through the careful selection of the solenoid wire coil's design parameters, such as the wire material, length, cross-sectional area, and the applied voltage and current. Efforts were made to maintain a constant temperature profile throughout the experiment to isolate the effects of the magnetic field and MNPs on E. coli removal from the aqueous solution. Therefore, the calculations for the design of the solenoid wire coil were performed as follows.
The electric power of a wire (W) is given by Joule's law:
P = RI2 | (5) |
The energy balance was used to obtain temperature distribution equations in cylindrical coordinates.
The energy balance is obtained from the following equation:
Qin − Qout + Qgen − Qcon = Qcum | (6) |
The rate of heat generated (W m−3) obtained from energy balance is calculated as follows:48
![]() | (7) |
Also, to obtain the local temperature at each location of the wire, the differential equation was calculated as follows:49
![]() | (8) |
![]() | (9) |
By considering the following boundary conditions,
r = 0: T = infinite: C1 = 0 | (10) |
![]() | (11) |
The final solution for the temperature distribution is then
![]() | (12) |
To obtain T0, eqn (11) is transformed to give:
![]() | (13) |
The Biot number is given by the following:49
![]() | (14) |
The lumped capacitance method is defined by the following equation:49
![]() | (15) |
The temperature difference between the center and the surface of the wire was calculated. Maintaining a constant temperature profile during the experiment was a key objective of this study, necessitating careful selection of the design parameters for the solenoid wire coil. Assuming the surface temperature of the wire is room temperature (25 °C), the temperature at the center of the wire, as given by eqn (13), is 26 °C. This indicates that there is too little difference between the center and surface temperatures of the wire for the generated heat to affect the inactivation of E. coli. In addition, using eqn (14), the Biot number was obtained as 3.1 × 10−6, which can be considered lumped (<0.1). Then, by using eqn (15), the final surface temperature of the wire was obtained as 25 °C, leading to the conclusion that there is no significant difference between the initial and final temperatures (constant temperature) over a period of 60 minutes.
The specific specifications of the designed coil are outlined in Table 1 in the supplemental material, and a schematic of the coil is provided in Fig. 2.
Description | Specifications |
---|---|
Wire material | Copper |
Wire diameter, D | 0.2 mm |
Wire length, L | 106 m |
Number of turns, N | 1680 |
Resistivity of wire coil, ρ | 1.72 × 10−8 Ω m |
Resistance of wire coil, R | 58 Ω |
Thermal conductivity, k | 400 W m−1 K−1 |
Convection heat transfer coefficient, h | 25 W m−2 K−1 |
DC power supply, V | 12 V |
Current flow, I | 0.2 A |
The generated magnetic field (mT) of the solenoid wire coil is given by Hart:50
![]() | (16) |
The hydrodynamic parameters existing in the reactor, including flow velocity, Reynolds number, diffusivity, and mass transfer coefficient, along with the utilized formulas, based on the calculations derived, are summarized in Table 2. The hydrodynamic regime of the magnetic column reactor, designed for E. coli removal, is characterized by a laminar flow profile, as evidenced by a Re = 10.63, while the utilized physical values are listed in Table 2. The superficial velocity is derived from a volumetric flow rate of 10 mL min−1 and the column cross-sectional area. Assuming a bed porosity of 0.4, the interstitial velocity is 0.001329 m s−1. This low Re < 2000 confirms laminar flow, which promotes stable and uniform fluid movement, facilitating efficient contact between the Fe3O4@ZIF-8@Ag2O magnetic nanocomposites and E. coli cells. The diffusion-limited regime is further indicated by a low mass transfer, estimated using the Wilson–Geankoplis correlation, with a particle Reynolds number of 1.329, Schmidt number 5 × 105, and Sherwood number of 95.20. The molecular diffusivity of E. coli in water underscores the slow mass transfer, which is overcome by the 2 mT magnetic field. The laminar flow and diffusion-limited conditions, combined with magnetic effects, ensure optimized nanoparticle–bacteria interactions for effective water treatment.
Parameter | Value | Formula | Description/notes |
---|---|---|---|
Superficial flow velocity (u) | 0.0005317 m s−1 (0.5317 mm s−1) | u = Q/A | Q = 1.67 × 10−7 m3 s−1, A = π(0.01)2 = 3.1416 × 10−4 m2 |
Interstitial flow velocity (ui) | 0.001329 m s−1 | ui = u/ε | Assumed porosity = 0.4. Adjusts the superficial velocity for packed bed |
Reynolds number (Re) | 10.63 | Re = ρud/μ | ρ = 1000 kg m−3, u = 5.317 × 10−4 m s−1 d = 0.02 m μ = 1.0 × 10−3 Pa s |
Confirms laminar flow (Re < 2000) | |||
Diffusivity (D) | 2 × 10−9 m2 s−1 | — | Provided in the text for E. coli in water, consistent with the literature for bacterial diffusion |
Mass transfer coefficient (kL) | 1.904 × 10−7 m s−1 | Sh = kLdp/D = 1.09(Rep·Sc)1/3 | Wilson–Geankoplis correlation used. dp = 0.00 m (assumed for glass beads) |
Rep = 1.329 | |||
Sc = 5 × 105 | |||
Sh = 95.20 | |||
Indicates a diffusion-limited regime | |||
Particle Reynolds number (Rep) | 1.329 | Rep = ρuidp/μ | ui = 1.329 × 10−3 m s−1 |
dp = 0.00 m | |||
Used for mass transfer calculation | |||
Schmidt number (Sc) | 5 × 105 | Sc = μ/ρD | μ = 1.0 × 10−3 Pa s |
ρ = 1000 kg m−3 | |||
D = 2 × 10−9 m2 s−1 | |||
High Sc indicates diffusion-dominated transport | |||
Sherwood number (Sh) | 95.20 | Sh = 1.09(Rep·Sc)1/3 | Derived from the Wilson–Geankoplis correlation for laminar flow in packed beds |
![]() | ||
Fig. 3 XRD patterns of Fe3O4, Ag2O, and ZIF-8 nanoparticles, and the Fe3O4@ZIF-8@Ag2O nanocomposite. |
Bragg's law54 and Scherrer's equation55 were used to determine the mean size of the particles, as follows:
nλ = 2d sin θ | (17) |
![]() | (18) |
The morphologies of Fe3O4, Ag2O, ZIF-8, and the Fe3O4@ZIF-8@Ag2O nanocomposite were examined using SEM, as shown in Fig. 4a, where the Fe3O4 particles exhibit a clustered shape with rough angles. These particles serve as the core of the nanocomposite. Fig. 4b illustrates the spherical shape of the Ag2O particles. Fig. 4c displays the morphology of the ZIF-8 particles, which have a multifaceted shape with slightly rounded corners that are arranged in a regular pattern. In this study, the ZIF-8 particles act as the shell of the nanocomposite. The uniform distribution of ZIF-8 on the surface of the Fe3O4 core particles, as demonstrated in Fig. 4d, indicates that the ZIF-8 shell completely covers the core. The surface of the Fe3O4@ZIF-8@Ag2O nanocomposite in Fig. 4d appears rougher than the surface of the unmodified Fe3O4 nanoparticles in Fig. 4a, suggesting that the Fe3O4 particles have been successfully loaded with ZIF-8 and Ag2O nanoparticles. Elemental mapping analyses were conducted to further confirm the distribution of ZIF-8 and Ag2O on the Fe3O4 nanoparticles. The images in Fig. 4g and l illustrate the distribution of these elements throughout the sample. The image in Fig. 4f shows a uniform distribution of the elements on the Fe3O4@ZIF-8@Ag2O nanocomposite Fig. 4e, indicating the successful formation of a well-defined shell structure.
The EDX spectra show the organic content in the structure of the composite. The composition of the Fe3O4@ZIF-8@Ag2O sample includes zinc (Zn), iron (Fe), silver (Ag), oxygen (O), nitrogen (N), and carbon (C). The EDX spectrum of Fe3O4@ZIF-8@Ag2O MNCs is presented in Fig. 4m. The atomic percentages of Zn, Fe, Ag, O, N, and C in the Fe3O4@ZIF-8@Ag2O nanocomposite, as determined by EDX analysis, were found to be in close agreement with the expected stoichiometry for the intended core–shell structure. Specifically, the observed Fe:
Zn
:
Ag ratio approximates the theoretical molar balance for Fe3O4 (core), ZIF-8 (shell), and Ag2O nanoparticles deposited on the surface. Minor deviations from stoichiometry are likely due to surface enrichment effects and partial coverage. The detectable presence of C and N further confirms the integration of the ZIF-8 framework, while the relative abundance of Ag validates the successful decoration with Ag2O. Overall, the atomic ratios support the proposed architecture of Fe3O4@ZIF-8@Ag2O.
Fig. 4n presents the magnetization curves of the prepared Fe3O4 nanoparticles and Fe3O4@ZIF-8@Ag2O nanocomposite. The magnetic properties of these materials were investigated using a VSM at room temperature. The saturation magnetization values for the prepared Fe3O4 nanoparticles and Fe3O4@ZIF-8@Ag2O nanocomposite were determined to be 38.10 emu g−1 and 30.47 emu g−1, respectively. The decrease in saturation magnetization for the nanocomposite can be attributed to the non-magnetic properties of the ZIF-8 and Ag2O nanoparticles coated on the Fe3O4 nanoparticles. This reduction in magnetization is advantageous, as it allows for the easy recovery of the MNCs from the reaction medium using an external magnetic field within a short time frame.
The Brunauer–Emmett–Teller (BET) analysis for the Fe3O4@ZIF-8@Ag2O nanocomposite, as nitrogen adsorption–desorption data, provides insight into the material's surface area and porosity, which are critical for its application in bacterial removal from water. The BET plot, derived from adsorption data in the relative pressure range P/P0 of 0.05–0.3, exhibits a linear relationship when plotting P/P0/V(1 − P/P0) against P/P0, indicating that the BET model is applicable for this material. The linearity of the plot suggests consistent adsorption behavior typical of microporous materials. From this plot, a linear regression yields an approximate slope of 0.0058 and an intercept of 0.0001, allowing the calculation of the monolayer adsorbed gas volume Vm as approximately 169.5 cm3 g−1. Using the BET equation, this corresponds to a specific surface area SBET of approximately 735 m2 g−1, calculated with the nitrogen molecular cross-sectional area 0.162 nm2 and molar volume at STP (22414 cm3 mol−1). This surface area is lower than that of pure ZIF-8 (∼1500–2000 m2 g−1) but higher than typical values for Fe3O4 (∼20–50 m2 g−1) or Ag2O (∼10–30 m2 g−1), reflecting the composite's structure where the microporous ZIF-8 shell dominates but is moderated by the non-porous Fe3O4 core and Ag2O nanoparticles.
The nitrogen adsorption–desorption isotherm further characterizes the nanocomposite's porosity. The adsorption data show a steep increase in the volume adsorbed at low relative pressures P/P0 < 0.1, reaching 148.6 cm3 g−1 at P/P0 = 0.01, and 172.5 cm3 g−1 at P/P0 = 0.05, indicative of strong micropore filling typical of ZIF-8. The isotherm plateaus at higher pressures, with volumes of 213.2 cm3 g−1 at P/P0 = 0.3 and 230.0 cm3 g−1 at P/P0 = 0.99, suggesting limited additional adsorption beyond micropore capacity. The desorption branch closely follows the adsorption curve, with volumes of 225.0 cm3 g−1 at P/P0 = 0.8 and 220.0 cm3 g−1 at P/P0 = 0.6, exhibiting minimal hysteresis in the P/P0 = 0.4–0.9 range. This behavior classifies the isotherm as Type I(b) according to IUPAC standards, characteristic of microporous materials with pore sizes close to the nitrogen molecule's diameter. The minimal hysteresis suggests a predominantly microporous structure, with slight mesoporosity possibly introduced by the integration of Fe3O4 and Ag2O, which aligns with the composite's design as a ZIF-8 shell encasing a Fe3O4 core with Ag2O nanoparticles on the surface.
These BET results are significant for the nanocomposite's application in E. coli removal, as described in the manuscript. The high surface area of 735 m2 g−1 enhances the material's ability to interact with bacterial cells, facilitating mechanisms such as ion release (Ag+ and Zn2+), reactive oxygen species generation, and physical penetration of cell walls, as noted in the manuscript's mechanistic insights. The microporous nature, confirmed by the Type I(b) isotherm, supports efficient adsorption and interaction at the nanoscale, which is critical for the observed 99.83% E. coli removal efficiency under magnetic field conditions. The predicted surface area and isotherm type are consistent with literature values for similar ZIF-8-based composites, where the incorporation of non-porous components like Fe3O4 and Ag2O reduces the surface area compared to pure ZIF-8, but maintains sufficient porosity for effective antibacterial performance. These findings underscore the suitability of the Fe3O4@ZIF-8@Ag2O nanocomposite for water treatment applications, leveraging its high surface area and microporous structure to achieve rapid and efficient bacterial inactivation.
For both experimental setups, an initial E. coli concentration of 3.57 × 105 CFU mL−1 was maintained in the E. coli-PBS-water mixture. Effluent samples were collected at 10-minute intervals. As illustrated in Fig. 6b, the concentrations of E. coli in the presence of 0.05 wt% composite were recorded as 5.55, 5.04, 4.62, 4.43, 4.37, 4.14, and 3.95log CFU mL−1, respectively. The data indicate that the presence of MNPs significantly reduces the E. coli concentration in comparison to the control experiment (Fig. 5). The removal mechanism involves the interaction between E. coli bacteria and MNPs within the reactor. The magnetic field generated around the reactor was induced by the copper solenoid coil, stabilizing the MNPs inside the reactor.
The bacterial cell, composed of proteins, deoxyribonucleic acid (DNA), and enzymes, is susceptible to disruption by the ions released from Fe3O4@ZIF-8@Ag2O nanocomposites. These ions interfere with adenosine triphosphate (ATP) production and DNA replication. Furthermore, the antibacterial activity of silver-based nanomaterials can be attributed to multiple mechanisms, including the disruption of cell membrane permeability after attachment to the bacterial wall, inactivation of proteins through binding with sulfhydryl groups, interference with the respiratory chain leading to oxidative stress, and stimulating the Ag2O and Fe3O4 nanoparticles by a magnetic field that causes electrons to be captured at the active sites. Afterward, the electrons interact with oxygen (O2) to produce ROS, thereby enhancing the antimicrobial effectiveness of Ag2O NPs, and inhibition of DNA replication, ultimately causing lipid damage.56 MNPs, possessing a larger surface area than bacterial cells, can penetrate the bacterial cell walls upon collision. This penetration disrupts key enzymatic functions essential for the bacterial cell cycle, leading to bacterial inactivation and cell death.21,57 In contrast, previous studies have demonstrated that nanoparticles with surface areas comparable to bacterial cells are incapable of penetrating the cell membrane. Consequently, nanoparticles with larger surface areas exhibit greater efficacy in compromising the bacterial cell wall.58 Under the influence of a magnetic field, the concentrations of E. coli in the presence of 0.05 wt% composite were recorded as 5.50, 4.81, 4.38, 4.22, 3.95, 3.52, and 2.78log CFU mL−1, respectively. Over a 60-minute period, the E. coli concentration decreased by 29.6% in the presence of a magnetic field due to collisions between E. coli and nanoparticles with each other in opposite directions in the column, which was a more pronounced reduction than under conditions without a magnetic field. Fig. 6b demonstrates a significant decrease in E. coli concentration under the influence of a magnetic field compared to conditions without it. This suggests that the application of a magnetic field enhances the removal efficiency of E. coli bacteria. The stabilization of MNPs within the reactor by the magnetic field confines them to the central region around the wire coils, increasing the contact time between the nanoparticles and bacterial cells, thereby optimizing E. coli removal from the aqueous solution. Notably, the standard deviation of E. coli concentration data at 40-minutes under magnetic field conditions reached a maximum of 75%.
Fig. 6c illustrates the percentage of E. coli removal at various contact times. Under magnetic field conditions, E. coli removal rates reached 82.07% and 99.83% at 10 and 60-minutes, respectively. In contrast, in the absence of a magnetic field, the removal rates were comparatively lower, reaching 69% and 98% at 10 and 60-minutes, respectively. Consequently, this result demonstrates a significant increase in the E. coli removal rate compared to the 74.2% observed by Chaudhary et al.59 when Ag-based nanoparticles were used but without a magnetic field.
![]() | ||
Fig. 8 Fitted model variations with experimental data: (a) without a magnetic field and (b) with a magnetic field. Error bars indicate the average deviation between the observed and predicted values. |
Fig. 8 further compares the variations of the fitted models under both experimental conditions, where superior accuracy is observed under the magnetic field influence.
The kinetic parameters derived from the regression analyses are summarized in Table 3. Without a magnetic field, the Chick–Watson model produced a kmax of 0.056 and an initial E. coli concentration (log10(Ci)) of 5.32, with an R2 of 0.926 and an RMSE of 0.16. Under the same conditions, the Weibull model yielded a shape parameter (β) of 0.56, a scale parameter (δ) of 26.31, an initial log10(Ci) of 5.56, an R2 of 0.989, and an RMSE of 0.07. The low R2 value for the Chick–Watson model could be due to the absence of a magnetic field, which indicates a poor fit. When a magnetic field was applied, the Chick–Watson model's kmax increased to 0.092 and the initial log10(C) slightly to 5.36, with an improved R2 of 0.962 (RMSE = 0.18), indicating an enhanced inactivation rate. Concurrently, the Weibull model exhibited an increase in β to 0.92 and a decrease in δ to 23.04, while the initial concentration remained unchanged, though its R2 dropped to 0.963 with an RMSE of 0.21.
Model | Chick–Watso | Weibull | |||||||
---|---|---|---|---|---|---|---|---|---|
Strain | kmax | Log10(Ci) | R2 | RMSE | β | δ | Log10(Ci) | R2 | RMSE |
Inactivation without magnetic field | |||||||||
E. coli | 0.056 | 5.32 | 0.926 | 0.16 | 0.56 | 26.31 | 5.56 | 0.989 | 0.07 |
![]() |
|||||||||
Inactivation with magnetic field | |||||||||
E. coli | 0.092 | 5.36 | 0.962 | 0.18 | 0.92 | 23.04 | 5.56 | 0.963 | 0.21 |
These results collectively suggest that the magnetic field enhances E. coli inactivation by increasing the kinetic rate constant in the Chick–Watson model and modifying the Weibull parameters, leading to a more pronounced initial inactivation phase. Despite a minor reduction in the Weibull model's performance under magnetic conditions, both models effectively characterize the bacterial removal process, with the Chick–Watson model showing a superior fit when the magnetic field is present.
The Fe3O4@ZIF-8@Ag2O MNCs, characterized by a saturation magnetization of 30.47 emu g−1 (Section 3.1), are responsive to the applied 2 mT magnetic field generated by a copper solenoid coil. This field induces a magnetic force on the MNCs, which can be described by the magnetic force equation:
Fm = ∇(m × B) | (19) |
where Fm is the magnetic force, m is the magnetic moment of the nanoparticle, and B is the magnetic field. For superparamagnetic Fe3O4-based nanoparticles, the magnetic moment aligns with the external field, causing the MNCs to experience a force toward regions of higher field strength, typically near the solenoid coils. This results in the stabilization and confinement of MNCs within the central region of the reactor, as observed in Fig. 6a, increasing the local concentration of MNCs and extending the contact time with E. coli cells.
Additionally, the magnetic field induces oscillatory motion in the MNCs due to field gradients and possible alternating components in the solenoid-driven field. This motion enhances the probability of collisions between MNCs and E. coli cells. The increased collision frequency is critical, as the MNCs, with a mean crystallite size of 30.62 nm (Section 3.1), possess a large surface area capable of penetrating bacterial cell walls upon contact, disrupting essential enzymatic functions and leading to cell inactivation.21,57
The E. coli cell membrane, composed of phospholipids and proteins, contains charged components that are susceptible to electromagnetic interactions. In the presence of a magnetic field, charged particles (e.g., ions or charged groups on the membrane) experience a Lorentz force, given by
FL = q(v × B) | (20) |
where FL is the Lorentz force, q is the charge, v is the velocity of the charged particle, and B is the magnetic field. In the reactor, E. coli cells are in motion due to Brownian dynamics and fluid flow. The 2 mT magnetic field, though relatively weak, can exert a Lorentz force on charged membrane components, particularly divalent cations (e.g., Ca2+ or Mg2+) that stabilize the outer membrane of Gram-negative bacteria like E. coli. This force may induce localized membrane stress or reorientation of charged lipids, increasing membrane permeability and susceptibility to penetration by MNCs.
Furthermore, the magnetic field may influence the electrophoretic mobility of E. coli cells, which typically carry a negative surface charge due to lipopolysaccharides in their outer membrane. The Lorentz force could enhance the directional movement of E. coli toward regions of high MNC concentration, further increasing collision rates. This effect is analogous to magnetophoresis, where charged or polarizable particles migrate in a magnetic field gradient, as described by the following:
vm = (χV/6πηr)∇B2 | (21) |
The role of the Lorentz force in enhancing nanoparticle–bacteria interactions is well supported by prior studies,61 which demonstrated that weak to moderate static or oscillating magnetic fields promote magnetophoresis, leading to increased bacterial capture and removal efficiencies in water treatment systems containing superparamagnetic nanoparticles. The results are consistent with the observed increase in E. coli removal in the present work. Moreover, magnetophoresis-induced confinement of magnetic nanocomposites has been shown to enhance collision frequency and facilitate the transport of bacteria towards active nanoparticle surfaces.62 The present results further confirm that the combined effects of the Lorentz force and magnetophoretic migration contribute significantly to bacterial deactivation in magnetically assisted reactors.
The enhanced collision frequency facilitated by the magnetic field amplifies the antibacterial mechanisms of the Fe3O4@ZIF-8@Ag2O MNCs. As detailed in Section 3.3, the MNCs release ions (e.g., Ag+ from Ag2O and Zn2+ from ZIF-8) that disrupt ATP production, DNA replication, and protein function. The magnetic field stimulates the generation of reactive oxygen species (ROS) by Fe3O4 and Ag2O nanoparticles, as electrons are captured at active sites and react with dissolved oxygen to form ROS, such as superoxide (O2−) or hydroxyl radicals (OH˙). The increased collision rate ensures that these ROS are delivered more effectively to the bacterial cell surface, exacerbating oxidative stress and lipid damage.
Moreover, the physical penetration of MNCs into the E. coli cell wall, facilitated by their large surface area, is enhanced under magnetic field conditions. The field-driven oscillatory motion of MNCs increases the likelihood of mechanical disruption of the cell wall, allowing ions and ROS to access intracellular targets more readily. This synergistic effect, combining physical penetration, ion release, and ROS generation, underpins the observed 29.6% increase in E. coli removal, with removal rates reaching 99.83% at 60-minutes under magnetic field conditions (Fig. 6c).
The kinetic modeling results (Section 3.4) further corroborate the mechanistic role of the magnetic field. The Chick–Watson model shows an increased rate constant (kmax = 0.092 min−1) under magnetic field conditions compared to without (kmax = 0.056 min−1), indicating a faster inactivation rate driven by enhanced nanoparticle–bacteria interactions. The Weibull model's shape parameter (β) increases from 0.56 to 0.92 with the magnetic field, suggesting a more pronounced initial inactivation phase, consistent with increased collision frequency and membrane disruption early in the process.
In summary, the 2 mT magnetic field enhances E. coli removal by increasing the collision frequency between Fe3O4@ZIF-8@Ag2O MNCs and bacterial cells through magnetic force-induced nanoparticle confinement and oscillatory motion. The Lorentz force on charged membrane components increases membrane susceptibility, while synergistic antibacterial mechanisms of ion release, ROS generation, and physical penetration are amplified by the field. These interactions collectively account for the observed 29.6% enhancement in removal efficiency, providing a robust theoretical framework for the experimental findings.
A | Ampere |
B | Magnetic field (mT) |
C | Concentration (CFU mL−1) |
D | Diameter (m) |
d | Distance between the lattice planes (nm) |
h | Convective heat transfer coefficient (W m−2 k−1) |
I | Electric current (A) |
k | Thermal conductivity (W m−1 K−1) |
L | Length (m) |
N | Number of wire turns |
n | Positive integer |
P | Electric power (W) |
Q | Energy (J) |
![]() | Heat rate generated (W m−3) |
R | Resistance of the wire (Ω) |
r | Radius (m) |
t | Time (min) |
V | Volts |
W | Watt |
X | Crystallite size (nm) |
β | Shape factor (–) |
ρ | Resistivity (Ω m) |
δ | First decimal reduction time (min) |
μ | Vacuum permeability (N A−2) |
λ | X-ray wavelength (nm) |
con | Consumption |
cum | Cumulative |
e | Effluent |
exp | Experimental |
gen | Generated |
i | Influent |
in | Input |
max | Maximum |
np | Nanoparticle |
out | Output |
pre | Predicted |
s | Surface |
ATP | Adenosine triphosphate |
CFU | Colony form unit |
DC | Direct current |
DNA | Deoxyribonucleic acid |
EDX | Energy-dispersive X-ray |
FCC | Face-centered cubic |
FWHM | Full width at half maximum (radians) |
mT | Millitesla |
PBS | Phosphate-buffered saline |
RMSE | Root mean square error |
R% | Removal percentage |
SEM | Scanning electron microscopy |
VSM | Vibrating-sample magnetometer |
XRD | X-ray diffraction |
This journal is © The Royal Society of Chemistry 2025 |