Next Article in Journal
An Evaluation of the Fire Safety of Waste Paper-Based Internal Finishing Materials Combined with Expandable Graphite According to Changes in Magnesium Hydroxide Content
Previous Article in Journal
A Review of Impedance Spectroscopy Technique: Applications, Modelling, and Case Study of Relative Humidity Sensors Development
Previous Article in Special Issue
The Concept of Using 2D Self-Assembly of Magnetic Nanoparticles for Bioassays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heat Generation and Diffusion in an Assembly of Magnetic Nanoparticles: Application to Magnetic Hyperthermia

by
Jean-Louis Déjardin
and
Hamid Kachkachi
*
Laboratoire PROMES-CNRS (UPR-8521), Université de Perpignan Via Domitia, Rambla de la Thermodynamique, Tecnosud, 66100 Perpignan, France
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(13), 5757; https://doi.org/10.3390/app14135757
Submission received: 29 April 2024 / Revised: 31 May 2024 / Accepted: 19 June 2024 / Published: 1 July 2024

Abstract

:
We investigate the thermal generation and transport properties of an assembly of magnetic nanoparticles embedded in a solid or fluid matrix, subjected to an AC magnetic field. For this purpose, we first build the heat equation for the assembly using the effective thermal transport coefficients obtained within the effective medium approach. In the present calculation, the SAR is obtained from the (linear) dynamic response of the assembly to the AC magnetic field. We numerically solve the extended heat equation and, as a preliminary study, we obtain the space-time profile of the temperature and total power absorbed by the system.

1. Introduction

Magnetic hyperthermia (MH) is a process by which magnetic nanoparticles (or nanomagnets, NMs), subjected to a time-dependent magnetic field (AFM), convert electromagnetic energy into heat. This process offers an advantage over the traditional methods of cancer treatment by irradiation [1,2,3,4,5,6,7,8], since it avoids the side effects of the latter by a directed and localized heating. Moreover, MH has been shown to kill cancer cells faster than the traditional methods and this reduces the therapy administration time [9]. In order to turn this technique into an efficient application, we still have to fully understand and control the mechanisms by which heat is generated and diffused in the assembly of NMs. The efficiency of MH is usually assessed through the specific absorption rate (SAR) [10,11,12,13,14,15,16,17,18]. Calorimetric methods provide another more precise means for characterizing MH and assessing its efficiency by studying the temperature rise after the start of heating, upon varying the AMF characteristics and the sample’s properties [19,20,21]. This technique is quite relevant in that it reduces the risk of heating the surrounding healthy cells. Today, a very lively debate is underway about the device-independence of these experiments, especially the measurement of the initial time slope of the temperature elevation [22].
In bulk homogeneous systems, heating mechanisms are well understood and the temperature rise is well described by the heat equation built on Fick’s and Fourier’s laws. However, at the nanoscale or in heterogeneous media, such as an assembly of NMs embedded in a solid matrix or immersed in a liquid, it is not obvious how the heat equation can be extended. Indeed, at reduced dimensions, even the definition of temperature itself becomes questionable. For heterogeneous media, several physical quantities, such as electrical and thermal conductivities and dielectric functions, have been calculated with the help of the effective medium approach [23,24,25,26,27,28,29,30,31,32,33,34]. This is a standard homogenization approach that leads to a macroscopic characterization of the medium with fewer parameters than those needed for a full description of the original system. For example, the whole micro-structure with the dielectric variation of a sample is reduced to an effective permittivity, which one then tries to relate to that of the inclusions and the matrix. The textbook [35] provides an extensive review of the topic and the various models and approaches that have been developed for tackling this issue.
In the present work, we follow this approach to provide expressions for the effective thermal conductivity and heat capacity of the assembly (or ferrofluid) of NMs. Then, replacing the latter by an effective medium with its effective transport coefficients, we write the corresponding extended heat equation with a source given by the heat generated by the NMs under the effect of the AMF. We then solve the extended heat equation using numerical tools and compute the space-time profile of the temperature elevation and total absorbed power for a typical assembly of magnetite nanoparticles.
We would like to emphasize that the aim of this work is not to present an extensive study of the results that can be inferred by varying all relevant physical parameters of the problem. It is not either an immediate application of MH in biological media and therapeutic treatment of cancer tumors. Such issues have been addressed by a few groups (see Refs. [36,37,38] and references therein) focusing on the bioheat equation and heat transfer in living tissues. Here, we propose a general formalism for studying heat generation and transport in a heterogeneous system and present a few results to illustrate the possible effects and observables that can be studied by such an approach. More precisely, we consider the system setup presented in Figure 1 below [39]. It is an assembly of nanomagnets (NMs), e.g., magnetite, embedded in a solid matrix (polymer, silica, etc) or floating in a fluid (ferrofluid), the base fluid (bf), which could be water. This system is in contact with its environment that may be the container of the sample (a tube, a box), air or a tissue. In the sequel, the subscript “ff” will stand for ferrofluid; “bf” for base fluid and “Fe” for iron (or magnetic substance of the inclusions). In the equations below, the quantities ρ , c v , κ , P and L refer to the whole sample (NM+host) and should then carry the subscript “ff”. However, to keep the equations simpler, we omit the subscript. We will introduce the density of the magnetic substance within the particles m Fe / V Fe = ρ Fe and the volume fraction of the ferrofluid ϕ V Fe / V ff .
In the general scheme for the measurement of the specific absorption rate (SAR, denoted as S ), the sample is initially set to the same temperature as its environment, T 0 . Then, the AC magnetic field (AMF) is switched on at t = 0 . The latter excites the NMs which then release heat within the sample, i.e., they heat up the matrix/fluid (ferrofluid). The heat thus generated then diffuses across the sample and the heat flux reaches the sample’s limits, or its interface with the environment, to which heat is transferred by conduction, convection or radiation. If we assume that the NMs are homogeneously distributed through the sample, the heat power by unit volume is P / V ff (that is in W / m 3 ).

2. Heat Equation

In general, the heat equation is an example of what is called IVP-BVP problem, i.e., a differential equation for a time- and space-dependent function f with both initial and boundary conditions, namely (e.g., for a solid medium of length L)
f t x , t κ x 2 f x , t = 0 , f x , 0 = T 0 , f 0 , t = α , f L , t = β .
α , β are given constants.
In our context, the space-time evolution of the temperature T r , t within a given medium, measured at position r and time t, may be described by the “extended” heat equation with source, namely
ρ c v T t r , t + ρ c v v r , t · T r , t κ 2 T r , t = P T r , t / V ff ,
where ρ kg / m 3 , c v J / K / kg , and κ W / K / m are, respectively, the density, the specific heat and the thermal conductivity of the medium (ferrofluid); v r , t is the local flow velocity (in m/s); and P W is the heat power. Note that the power P depends on the temperature T r , t . Indeed, in our context, P is generated by the NMs subjected to an AMF which drives its magnetization into an oscillatory motion with a characteristic time that depends on the temperature in the immediate vicinity of the NMs, namely the temperature T r , t . More precisely, T enters P through the AC susceptibility and the relaxation time of the NMs [see Section 2.3].
In Equation (2) one implicitly assumes that the sample is insulated at its interface with the outer environment (air in the room, other parts of a body, etc). In reality, this is not rigorously true and the sample does exchange heat with its environment at room temperature T 0 . One of the mechanisms for such an exchange is given by Newton’s law of cooling which states that the rate of temperature change as a result of the sample-environment exchange is (for the whole sample of volume V s )
C T t r , t = L T r , t T 0 ,
where L is the so called Newton (positive) coefficient measured in W/K and C J / K the heat capacity of the sample. See discussion in Ref. [39] and references therein.

2.1. Extended Heat Equation

As discussed in the introduction, an example of the systems targeted by the present study is sketched in Figure 1. In the effective media approach, this heterogeneous system, with parameters ρ NP , κ NP (for the nanoparticles or NMs) and ρ m , κ m (for the matrix), is replaced by a homogeneous effective system with the effective parameters ρ eff , κ eff (the subscript “eff” is dropped in the sequel), as sketched in Figure 2.
Equation (2) may be extended to include the contribution from the interface, leading to the following more general equation (here L is measured in W/K/m3)
ρ c v T t r , t + ρ c v v r , t · T r , t κ 2 T r , t = P V ff L T r , t T 0 .
Dividing Equation (4) by T 0 , we rewrite it in terms of θ r , t T r , t T 0 / T 0 , the relative temperature elevation, as follows:
ρ c v θ t r , t + ρ c v v r , t · θ r , t κ 2 θ r , t = q θ ,
where we have defined the source as a function of θ :
q θ 1 T 0 P θ V ff L θ .
Next, dividing by L, we note that ρ c v / L has the dimensions of time and accordingly introduce the characteristic time
t s ρ c v L .
Then, we rewrite Equation (5) as follows:
θ τ r , τ + v ˜ r , τ · θ r , τ κ ˜ 2 θ r , τ = q ˜ θ .
Here, we also introduce the (dimensionless) time τ = t / t s and the new quantities
v ˜ t s v , κ ˜ κ L , q ˜ θ q L = θ + 1 L T 0 P θ V ff .
We may rewrite q ˜ using the SAR S = P / m Fe (in W/kg) as follows:
q ˜ θ = θ + S L T 0 ρ Fe .
The quantity S L T 0 ρ Fe ϕ is the dimensionless SAR Ξ θ introduced in Ref. [39].
Consequently, we have the final (dimensionless) heat equation
θ τ r , τ + v ˜ r , τ · θ r , τ κ ˜ 2 θ r , τ = θ + Ξ θ .

2.2. Physical Parameters

As discussed earlier, we may consider a system composed of a fluid (here, water) and magnetic inclusions (maghemite or magnetite nanoparticles). An important relevant quantity is then the mass of the whole magnetic material introduced into the fluid, which we denote by m Fe . In practice, we need the SAR S which is the power P divided by the mass of the magnetic substance m Fe . In addition, we note that since c v , ρ , κ and L refer to the ferrofluid, they depend on the volume fraction ϕ of the inclusions, namely the number of NMs introduced in the ferrofluid volume ( V ff ), in addition to the corresponding parameters of the nanofluid components.
  • Specific heat: The sample (ferrofluid) specific heat is given by
    c v ϕ = ϕ ρ Fe c v Fe + 1 ϕ ρ bf c v bf ϕ ρ Fe + 1 ϕ ρ bf ,
    with
    ρ ϕ = ϕ ρ Fe + 1 ϕ ρ bf .
  • Thermal conductivity: There are several models for κ ϕ of a mixture based on the effective-medium approach from which we choose that of Yu and Choi [40]
    κ ϕ = κ m κ pm + 2 + 2 ϕ κ pm 1 1 + β 3 κ pm + 2 ϕ κ pm 1 1 + β 3 ,
    with β being the ratio of the thickness of the shell surrounding the nanoparticle to the radius of the latter and κ pm κ Fe / κ m the relative thermal conductivity.
  • Power vs. SAR: In Ref. [39] we already saw that the power P divided by the mass of the magnetic substance ( m Fe ) yields the SAR S in W/kg. So
    P m Fe = S .
    See discussion before Equation (11).

2.3. Specific Absorption Rate

From previous works [16,39], using the theory of linear magnetic response, the SAR can be written as
Ξ = Ξ 0 h , ω Φ θ
where
Ξ 0 h μ 0 h 2 m 2 6 k B l ff T 0 2 ω , Φ θ η θ 1 + θ 1 + η 2 θ .
Here
η θ = π 2 ϖ 0 1 + θ e σ 0 1 + θ σ 0 , ϖ 0 ω τ 0 , σ 0 K V k B T 0 .
We also introduce the following quantity
η 0 ω = π 2 ϖ 0 e σ 0 σ 0 = η T 0 , ω .
To plot the SAR as a function of temperature, we use the experimental data from Ref. [21] for magnetite ferrofluid: nanoparticles of iron oxide with a concentration of 8.6 mg/cm3, a diameter of 9 nm, anisotropy constant K = 3 × 10 4 J / m 3 , a saturation magnetization M s = 480 × 10 3 A / m , and a density ρ = 5.2 g / cm 3 . T 0 = 318 K , frequency f = ω / 2 π = 194 × 10 3 Hz , and magnetic field h = 38.2 kA / m . Finally, from Ref. [39] we have τ 0 = 9.5 × 10 10 s and l ff = 0.330 W / K / g . The plot of the corresponding SAR is shown in Figure 3 for an assembly with lognormal size distribution with three values of the standard deviation and the same mean diameter.
As discussed in Ref. [39], θ is small for prototypical ferrofluids and, as such, we may expand Ξ θ , ϕ , appearing in Equation (11), in powers of θ and write to first order: Ξ Ξ 0 + Ξ 1 θ . Then, the right-hand side of Equation (11) becomes Ξ 0 + Ξ 1 1 θ . Indeed, we expand the function Φ in powers of θ :
Φ θ Φ 0 + Φ 1 θ + O θ 3
with [ Φ n ω = Φ n ω , T 0 ]
Φ 0 = η 0 1 + η 0 2 , Φ 1 = η 0 2 1 + η 0 2 2 1 + 2 σ 0 + 3 2 σ 0 η 0 2 .
Therefore, the SAR expansion reads
Ξ = Ξ 0 h , ω η 0 1 + η 0 2 Ξ 0 h , ω η 0 2 1 + η 0 2 2 1 + 2 σ 0 + 3 2 σ 0 η 0 2 × θ
and thereby the quantities Ξ 0 and Ξ 1 introduced above are explicitly given by
Ξ 0 = Ξ 0 h , ω η 0 1 + η 0 2 = 1 l ff T 0 μ 0 h 2 m 2 6 k B T 0 ω η 0 1 + η 0 2 , Ξ 1 = Ξ 0 h , ω , T 0 1 + 2 σ 0 + 3 2 σ 0 η 0 2 2 1 + η 0 2 .
Note that Ξ 0 is the (dimensionless) SAR evaluated at T 0 , namely Ξ h , ω , T 0 .
Therefore, we have the final heat equation
τ θ r , t + v ˜ ϕ · θ r , t κ ˜ ϕ 2 θ r , t = Ξ h , ω , T 0 + Ξ 1 1 θ r , t
with
v ˜ ϕ = v L ϕ ρ Fe c v Fe + 1 ϕ ρ bf c v bf , κ ˜ ϕ = κ m L κ pm + 2 + 2 ϕ κ pm 1 1 + β 3 κ pm + 2 ϕ κ pm 1 1 + β 3 , Ξ 0 h , ω , T 0 = μ 0 h 2 m 2 6 k B l ff T 0 2 ω × η 0 1 + η 0 2 , Ξ 1 = Ξ 0 h , ω , T 0 1 + 2 σ 0 + 3 2 σ 0 η 0 2 2 1 + η 0 2 .
In the present study we ignore the shell, i.e., β = 0 . We also assume that the particles are fixed within the host and thereby ignore convection effects in this study.

3. Results and Discussion

We have used numerical tools to solve Equation (11) or (24) with the physical parameters of magnetite. For the dependent variable θ r , t , i.e., the relative temperature elevation, we use the initial/boundary conditions,
θ r , 0 = 0 , θ 0 , t = θ L , t = 0 .
A more realistic situation of three dimensions is supposed to mimic the fact that the ferrofluid is embedded in a medium that is supposed to be kept at a constant temperature T 0 (∼ 42 45   ° C ); recall that θ = T T 0 / T 0 . This is, for instance, the situation of a suspension of magnetic nanoparticles injected in a tumor, where the boundary conditions above prevent overheating of the surrounding area.
In this preliminary study, we only consider the case of a two-dimensional assembly of NMs, as shown in Figure 4, which may be either organized or randomly distributed in the matrix. In the present work, this does not matter since we do not include dipolar interactions between the NMs. The effect of the latter on the SAR was studied in Refs. [16,17,18] and references therein.
More precisely, we consider assemblies of magnetite particles as studied experimentally in Ref. [21] and, later, theoretically in Ref. [39]. In Table 1, we collect all the relevant data used in the following figures.
The intensity of the AC magnetic field is 38.2 kA / m and its frequency is 123 kHz . Newton’s coefficient was obtained in Ref. [39] by fitting the experimental data from Ref. [21]: l ff = 330 W / K / kg . The NMs are lognormal distributed in size with mean diameter D m = 9 nm and standard deviation δ = 0.25 . Shah et al. [21] used 1.2 mL of magnetite ferrofluid (magnetite in water) and a concentration ρ Fe 8.6 mg / mL 8.6 kg / m 3 of magnetite in the ferrofluid. The rest of the data related to the magnetization dynamics are given in Section 2.2. of Ref. [39]. In particular, the relaxation time related to the Brownian motion of NMs in the fluid is given and used to define the effective relaxation time.
As discussed earlier, heat Equation (11) can be solved with free boundary conditions or with boundaries at fixed temperatures. The first case would apply in a situation where, for instance, we were interested in the spatial extent of heating. The second case with fixed boundaries would apply to a situation where the assembly is confined in some space brought into contact with a surrounding that is kept at a fixed temperature.
Next, the heat equation can also be applied to an assembly of NMs organized or not on a solid matrix, as shown in Figure 4. In the following, all results will be shown for an assembly of NMs that are randomly distributed in space.
As already stressed in the Introduction, this work is not an exhaustive study of the effect of the numerous physical parameters of the problem at hand, namely those related with NM assembly (size, shape, distributions, and concentration) and external excitation (amplitude and frequency). The main objective instead is to propose a formalism and a toolbox for investigating the heat generation and transport in assemblies of NMs with the possibility to vary all the relevant parameters. In the present study, we only show a sample of results to illustrate how our formalism can be used and what physical observables can be obtained. By way of example, in the following, we discuss the calculation of the total heat injected into the assembly at time τ , which may be defined as ( V s is the volume of the sample):
Δ Q τ Q t Q 0 = k B T 0 × 1 V s V s d r θ r , t .
In Figure 5, we compare the two situations with fixed boundary conditions (BCs) and free BC. We show both the 2 D (relative) temperature field and Δ Q . The plot of the latter clearly shows that the maximum heat is reached much more quickly in an assembly with fixed temperature T 0 (i.e., θ = 0 ) at the boundaries. As is corroborated by the 2 D temperature field, in the case with fixed BC, the heat cannot propagate freely but is confined in a smaller region. However, this depends on the heat capacity and thermal conductivity of the system, together with other physical parameters such as the concentration and spatial arrangements of the NMs. This is shown in the sequel.
Indeed, in Figure 6, we show the results of Δ Q t for a monodisperse assembly of NMs with a diameter D and two size-distributed assemblies with mean diameter smaller or larger than D. It is seen that globally the generated heat increases with the mean diameter, and this is due to the fact that the SAR (the heat source) also increases. The plots also show that the initial slope of Δ Q t depends on the size distribution of the assembly, and this is related to the fact that the diffusion of heat depends on the size and spatial distributions of the assembly.
In Figure 7, we compare assemblies with different values of the thermal conductivity of the host ( κ m ) (see Equations (14) and (25)). The first observation for such randomly distributed assemblies is that the result strongly depends on the time interval, and for long times we see that in a host with higher conductivity, the maximum heat is reached much faster.

4. Conclusions

In this work, we have presented a preliminary study of heat generation and transport in an assembly of magnetic nanoparticles subjected to a time-dependent magnetic field. For this purpose we have proposed an extended heat equation based on the effective medium approach which allows us to obtain the effective thermal transport coefficients for a heterogeneous medium such as an assembly of nanomagnets. Then, we numerically solved the extended heat equation and discussed the space-time profile of the temperature elevation for boundaries with fixed temperature or with free boundaries.
The present methodology may form the basis for studying the heat generation and transport of an assembly of nanomagnets. It makes it possible to compute, in a self-contained manner, the specific absorption rate and the space-time profile of the temperature elevation as functions of the various magnetic materials parameters, such as the concentration of the assembly, its spatial organization and size distribution, the convection coefficient, and the characteristics of the external excitation.
A comparison with experiments and a fit for the relevant physical parameters were performed in Ref. [39] for the time profile of the temperature elevation in magnetic ferrofluids. Such a comparison for the spatial profile requires measurements of the local temperature within the assembly [41,42,43].

Author Contributions

Conceptualization, H.K.; Methodology, H.K.; Formal analysis, J.-L.D.; Investigation, J.-L.D. and H.K.; Writing—review & editing, J.-L.D. and H.K.; Supervision, H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been partly funded by ANR under the contract ANR-21-CE09-0043-01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data and results included in this work are available on demand.

Acknowledgments

The authors acknowledge financial support from ANR-21-CE09-0043-01.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Overgaard, J. (Ed.) History and heritage: An introduction. In Hyperthermic Oncology; Taylor and Francis: London, UK, 1985; Volume 2, pp. 8–9. [Google Scholar]
  2. Van der Zee, J. Heating the patient: A promising approach? Ann. Oncol. 2002, 13, 1173. [Google Scholar] [CrossRef] [PubMed]
  3. Johannsen, M.; Gneveckow, U.; Eckelt, L.; Feussner, A.; Waldöfner, N.; Scholz, R.; Deger, S.; Wust, W.; Loening, S.A.; Jordan, A. Clinical hyperthermia of prostate cancer using magnetic nanoparticles: Presentation of a new interstitial technique. Int. J. Hyperth. 2005, 21, 637. [Google Scholar] [CrossRef] [PubMed]
  4. Skumiel, A.; Jósefczak, A.; Timko, M.; P Kopčanský, P.; Herchl, F.; Koneracká, M.; Tomašovičová, N. Tomašovičová. Int. J. Thermodyn. 2007, 28, 1461. [Google Scholar] [CrossRef]
  5. Murase, K.; Takata, H.; Takeuchi, Y.; Saito, S. Control of the Temperature Rise in Magnetic Hyperthermia with Use of an External Static Magnetic Field. Phys. Med. 2013, 29, 624. [Google Scholar] [CrossRef] [PubMed]
  6. Streffer, C.; van Beuningen, D. The biological basis for tumor therapy by hyperthermia and radiation. In Hyperthermia and the Therapy of Malignant Tumors; Streffer, C., Ed.; Springer: Berlin/Heidelberg, Germany, 1987; pp. 24–70. [Google Scholar]
  7. Haider, Q.; Hussain, A.; Rehman, A.; Ashour, A.; Althobaiti, A. Mass and Heat Transport Assessment and Nanomaterial Liquid Flowing on a Rotating Cone: A Numerical Computing Approach. Nanomaterials 2022, 12, 1700. [Google Scholar] [CrossRef] [PubMed]
  8. Siddique, S.; Chow, J.C.L. Recent Advances in Functionalized Nanoparticles in Cancer Theranostics. Nanomaterials 2022, 12, 2826. [Google Scholar] [CrossRef] [PubMed]
  9. Rodríguez-Luccioni, H.L.; Latorre-Esteves, M.M.; Méndez-Vega, J.J.; Soto, O.O.; Rodríguez, A.R.; Rinaldi, C.C.; Torres-Lugo, M.M. Enhanced reduction in cell viability by hyperthermia induced by magnetic nanoparticles. Int. J. Nanomed. 2011, 6, 373. [Google Scholar]
  10. Carrey, J.; Mehdaoui, B.; Respaud, M. Simple models for dynamic hysteresis loop calculations of magnetic single-domain nanoparticles: Application to magnetic hyperthermia optimization. J. Appl. Phys. 2011, 109, 083921. [Google Scholar] [CrossRef]
  11. Mehdaoui, B.; Meffre, A.; Carrey, J.; Lachaize, S.; Lacroix, L.-M.; Gougeon, M.; Chaudret, B.; Respaud, M. Optimal Size of Nanoparticles for Magnetic Hyperthermia: A Combined Theoretical and Experimental Study. Adv. Funct. Mater. 2011, 21, 4573. [Google Scholar] [CrossRef]
  12. Martinez-Boubeta, C.; Simeonidis, K.; Makridis, A.; Angelakeris, M.; Iglesias, O.; Guardia, P.; Cabot, A.; Yedra, L.; Estradé, S.; Peiró, F.; et al. Learning from Nature to Improve the Heat Generation of Iron-Oxide Nanoparticles for Magnetic Hyperthermia Applications. Sci. Rep. 2013, 3, 1652. [Google Scholar] [CrossRef]
  13. Conde-Leboran, I.; Baldomir, D.; Martinez-Boubeta, C.; Chubykalo-Fesenko, O.; del Morales, M.P.; Salas, G.; Cabrera, D.; Camarero, J.; Teran, F.J.; Serantes, D. A Single Picture Explains Diversity of Hyperthermia Response of Magnetic Nanoparticles. J. Phys. Chem. C 2015, 119, 15698. [Google Scholar] [CrossRef]
  14. Kostopoulou, A.; Lappas, A. Colloidal magnetic nanocrystal clusters: Variable length-scale interaction mechanisms, synergetic functionalities and technological advantages. Nanotechnol. Rev. 2015, 4, 595. [Google Scholar] [CrossRef]
  15. Lacroix, L.-M.; Malaki, R.B.; Carrey, J.; Lachaize, S.; Respaud, M.; Goya, G.F.; Chaudret, B. Magnetic hyperthermia in single-domain monodisperse FeCo nanoparticles: Evidences for Stoner–Wohlfarth behavior and large losses. J. Appl. Phys. 2009, 105, 023911. [Google Scholar] [CrossRef]
  16. Déjardin, J.-L.; Vernay, F.; Respaud, M.; Kachkachi, H. Effect of dipolar interactions and DC magnetic field on the specific absorption rate of an array of magnetic nanoparticles. J. Appl. Phys. 2017, 121, 203903. [Google Scholar] [CrossRef]
  17. Déjardin, J.-L.; Vernay, F.; Kachkachi, H. Specific absorption rate of magnetic nanoparticles: Nonlinear AC susceptibility. J. Appl. Phys. 2020, 128, 143901. [Google Scholar] [CrossRef]
  18. Déjardin, J.-L.; Vernay, F.; Kachkachi, H. Efficiency of energy dissipation in nanomagnets: A theoretical study of ac susceptibility. In Magnetic Nanoparticles in Human Health and Medicine; Rai, M., Caizer, C., Eds.; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  19. Murase, K.; Oonoki, J.; Takata, H.; Song, R.; Angraini, A.; Ausanai, P.; Matsushita, T. Simulation and experimental studies on magnetic hyperthermia with use of superparamagnetic iron oxide nanoparticles. Radiol. Phys. Technol. 2011, 4, 194. [Google Scholar] [CrossRef] [PubMed]
  20. Andreu, I.; Natividad, E. Accuracy of available methods for quantifying the heat power generation of nanoparticles for magnetic hyperthermia. Int. J. Hyperth. 2013, 29, 739. [Google Scholar] [CrossRef]
  21. Shah, R.R.; Davis, T.P.; Glover, A.L.; Nikles, D.E.; Brazel, C.S. Impact of magnetic field parameters and iron oxide nanoparticle properties on heat generation for use in magnetic hyperthermia. J. Magn. a Magn. Mater. 2015, 387, 96. [Google Scholar] [CrossRef] [PubMed]
  22. Ruta, S.; Fernández-Afonso, Y.; Rannala, S.E.; Morales, M.P.; Veintemillas-Verdaguer, S.; Jones, C.; Gutiérre, L.; Chantrell, R.W.; Serantes, D. A device-independent approach to evaluate the heating performance during magnetic hyperthermia experiments: Peak analysis and zigzag protocol. arXiv 2023, arXiv:2307.11521. [Google Scholar]
  23. Maxwell, J.C. A Treatise on Electricity and Magnetism; Reprinted 1998; Oxford University Press: Oxford, UK, 1891. [Google Scholar]
  24. Rayleigh, J.W.S. On the Instability of a Cylinder of Viscous Liquid under Capillary Force. Phil. Mag. 1892, 34, 431. [Google Scholar] [CrossRef]
  25. Garnett, J.C.M. Colours in Metal Glasses and in Metallic Films. Phil. Trans. R Soc. Lond. 1904, 203, 385. [Google Scholar] [CrossRef]
  26. Xuan, Y.; Li, Q. Investigation on convective heat transfer and flow features of nanofluids. J. Heat Transf. 2003, 125, 151. [Google Scholar] [CrossRef]
  27. Xuan, Y.; Li, Q.; Zhang, X.; Fujii, M. Stochastic Thermal Transport of Nanoparticle Suspensions. J. Appl. Phys. 2006, 100, 855. [Google Scholar] [CrossRef]
  28. Bird, R.B.; Stewart, W.E.; Lightfoot, E.N. Transport Phenomena; John Wiley & Sons: Hoboken, NJ, USA, 2007. [Google Scholar]
  29. Kakac, S.; Cao, L. Analysis of convective two-phase flow instabilities in vertical and horizontal in-tube boiling systems. Int. J. Heat Mass Transfer. 2009, 52, 3984. [Google Scholar] [CrossRef]
  30. Desai, T.G. Thermal transport in nanoclusters. J. Appl. Phys. 2011, 98, 193107. [Google Scholar] [CrossRef]
  31. Goharkhah, M.; Salarian, A.; Ashjaee, M.; Shahabadi, M. Convective heat transfer characteristics of magnetite nanofluid under the influence of constant and alternating magnetic field. Powder Tech. 2015, 274, 258. [Google Scholar] [CrossRef]
  32. Karimi, A.; Goharkhah, M.; Ashjaee, M.; Shafii, M. Thermal Conductivity of and Fe3O4 Magnetic Nanofluids Under the Influence of Magnetic Field. Int. J. Thermophys. 2015, 36, 2720. [Google Scholar] [CrossRef]
  33. Zouli, N. Enhancement of Heat Transfer Coefficient using Fe2O3-Water Nanofluids. Mater. Energy 2017, 2, 247. [Google Scholar]
  34. Maurice, M.; Barros, N.; Kachkachi, H. Beyond the Maxwell Garnett approximation for interacting plasmonic nanoparticles: An analytical and numerical study. J. Appl. Phys. 2023, 134, 094303. [Google Scholar] [CrossRef]
  35. Sihvola, A.H. Electromagnetic Mixing Formulas and Applications; IET Electromagnetic Waves Series; IET: London, UK, 1999; Volume 47. [Google Scholar]
  36. Andreozzi, A.; Brunese, L.; Iasiello, M.; Tucci, C.; Vanoli, G.P. Modeling heat transfer in tumors: A review of thermal therapies. Ann. Biomed. Eng. 2019, 47, 676. [Google Scholar] [CrossRef]
  37. Astefanoaei, I.; Stancu, A. A computational study of the bioheat transfer in magnetic hyperthermia cancer therapy. J. Appl. Phys. 2019, 125, 194701. [Google Scholar] [CrossRef]
  38. Rytov, R.A.; Bautin, V.A.; Usov, N.A. Towards optimal thermal distribution in magnetic hyperthermia. Sci. Rep. 2022, 12, 1. [Google Scholar]
  39. Déjardin, J.-L.; Kachkachi, H. Time profile of temperature rise in assemblies of nanomagnets. J. Magn. Magn. Mater. 2022, 556, 169354. [Google Scholar] [CrossRef]
  40. Yau, W.; Choi, S. The Role of Interfacial Layers in the Enhanced Thermal Conductivity of Nanofluids: A Renovated Maxwell Model. J. Nanopart. Res. 2003, 5, 167. [Google Scholar] [CrossRef]
  41. Jin, C.Y.; Li, Z.; Williams, R.S.; Lee, K.C.; Park, I. Localized temperature and chemical reaction control in nanoscale space by nanowire array. Nano Lett. 2011, 11, 4818–4825. [Google Scholar] [CrossRef] [PubMed]
  42. Kawamoto, N.; Wang, M.-S.; Wei, X.; Tang, D.-M.; Murakami, Y.; Shindo, D.; Mitome, M.; Golberg, D. Local temperature measurements on nanoscale materials using a movable nanothermocouple assembled in a transmission electron microscope. Nanotechnology 2011, 22, 485707. [Google Scholar] [CrossRef]
  43. Balois, M.V.; Hayazawa, N.; Catalan, F.C.; Kawata, S.; Yano, T.A.; Hayashi, T. Tip-enhanced THz Raman spectroscopy for local temperature determination at the nanoscale. Anal. Bioanal. Chem. 2015, 407, 8205–8213. [Google Scholar] [CrossRef]
Figure 1. Sketch of the model system: a sample composed of nanomagnets (NMs) as heat sources, and the embedding fluid (base fluid), coupled to its environment (air). Source: Ref. [39].
Figure 1. Sketch of the model system: a sample composed of nanomagnets (NMs) as heat sources, and the embedding fluid (base fluid), coupled to its environment (air). Source: Ref. [39].
Applsci 14 05757 g001
Figure 2. Effective medium approach: replacing the heterogeneous system with parameters ρ NP , κ NP (for the nanoparticles) and ρ m , κ m (for the matrix) by a homogeneous system with the effective parameters ρ eff , κ eff .
Figure 2. Effective medium approach: replacing the heterogeneous system with parameters ρ NP , κ NP (for the nanoparticles) and ρ m , κ m (for the matrix) by a homogeneous system with the effective parameters ρ eff , κ eff .
Applsci 14 05757 g002
Figure 3. Specific absorption rate for magnetite ferrofluids as a function of temperature and varying standard deviation δ .
Figure 3. Specific absorption rate for magnetite ferrofluids as a function of temperature and varying standard deviation δ .
Applsci 14 05757 g003
Figure 4. Assembly of organized or randomly distributed nanoparticles.
Figure 4. Assembly of organized or randomly distributed nanoparticles.
Applsci 14 05757 g004
Figure 5. Left: 2D temperature field for both fixed BC and free BC. Right: Time profile of the total heat for fixed boundaries and free boundaries (right).
Figure 5. Left: 2D temperature field for both fixed BC and free BC. Right: Time profile of the total heat for fixed boundaries and free boundaries (right).
Applsci 14 05757 g005
Figure 6. Upper panel: 2D temperature field for three assemblies (see text). Lower graph: Time profile of the total heat generated in these assemblies.
Figure 6. Upper panel: 2D temperature field for three assemblies (see text). Lower graph: Time profile of the total heat generated in these assemblies.
Applsci 14 05757 g006
Figure 7. Time profile of the total heat for three different values of the thermal conductivity of the host ( κ m ).
Figure 7. Time profile of the total heat for three different values of the thermal conductivity of the host ( κ m ).
Applsci 14 05757 g007
Table 1. Physical parameters from Refs. [21,39].
Table 1. Physical parameters from Refs. [21,39].
Substance c v J / K / kg κ W / K / m ρ kg / m 3 m ( kg ) V ( m 3 ) K ( J / m 3 ) M s A / m
Magnetite6195.15200 1.032 × 10 5 1.985 × 10 9 3.0 × 10 4 4.8 × 10 5
Water41810.5981000 1.198 × 10 3 1.198 × 10 6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Déjardin, J.-L.; Kachkachi, H. Heat Generation and Diffusion in an Assembly of Magnetic Nanoparticles: Application to Magnetic Hyperthermia. Appl. Sci. 2024, 14, 5757. https://doi.org/10.3390/app14135757

AMA Style

Déjardin J-L, Kachkachi H. Heat Generation and Diffusion in an Assembly of Magnetic Nanoparticles: Application to Magnetic Hyperthermia. Applied Sciences. 2024; 14(13):5757. https://doi.org/10.3390/app14135757

Chicago/Turabian Style

Déjardin, Jean-Louis, and Hamid Kachkachi. 2024. "Heat Generation and Diffusion in an Assembly of Magnetic Nanoparticles: Application to Magnetic Hyperthermia" Applied Sciences 14, no. 13: 5757. https://doi.org/10.3390/app14135757

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop