Next Article in Journal
A Combined Multi-Classification Network Intrusion Detection System Based on Feature Selection and Neural Network Improvement
Previous Article in Journal
Radish Growth Stage Recognition Based on GAN and Deep Transfer Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biothermal Heating on Human Skin by Millimeter and Sub-Terahertz Waves in Outdoor Environment—A Theoretical Study

1
School of Integrated Circuits and Electronics, Beijing Institute of Technology, Beijing 100081, China
2
Tangshan Research Institute, Beijing Institute of Technology, Tangshan 063099, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(14), 8305; https://doi.org/10.3390/app13148305
Submission received: 23 May 2023 / Revised: 12 July 2023 / Accepted: 17 July 2023 / Published: 18 July 2023

Abstract

:
The frequency band in the millimeter-wave (MMW) and sub-terahertz (sub-THz) range has shown great potential in mobile communication technology due to the advantages of ultra-large bandwidth and ultra-high data rates. Based on the increasing research activities on MMW/sub-THz waves, biological safety at relevant frequencies must be explored, especially when high-power illumination occurs. Here, its non-ionizing nature plays a vital role, which makes it safe for humans at low illumination powers. However, under high power, the biothermal heating on the skin surface is still a main concern, and lots of research has been conducted in a laboratory. In this article, we analyze the thermal heating effect of human skin in outdoor environments, where atmospheric conditions can significantly impact the propagation of MMW/sub-THz waves. Our analysis is based on rat skin, which has a similar structure to human skin. A theoretical model combining Pennes’ bioheat transfer equation (BHTE), the ITU model, and the Mie scattering theory is developed. Good agreement between calculation results and measured data confirms the efficiency of this model. The influence of rainfall rate, humidity, operating frequency, illumination time, power density, and propagation distance is presented and discussed.

1. Introduction

As the outer tissue of the human body, the skin acts as an interface with the environment and thus is the first to be exposed to electromagnetic radiation, especially at high illumination powers. When the irradiation power density is higher than 10 mW/cm2, the thermal effect is mainly studied, which is directly manifested by increasing skin temperature and damage to biological tissues [1]. There have been several safety guidelines/standards set by the International Commission on Non-Ionizing Radiation Protection (ICNIRP) and the IEEE International Committee on Electromagnetic Safety (ICES) Technical Committee (TC) 95 [2,3,4] to protect humans from excessive exposure. Reference levels of 20 mW/cm2 at 6 GHz to 10 mW/cm2 at 300 GHz have been derived for local body exposure based on laboratory measurements or simulations [5,6,7]. However, further research on skin response in the atmospheric environment is required for outdoor applications of MMW/sub-THz techniques, such as remote security inspection, high-speed data coverage, and active denial system.
MMW/sub-THz waves suffer from high sensitivity to outdoor weather conditions [8,9,10,11,12,13,14,15]. The path availability and channel performance are reduced even in clear weather, as atmosphere gases and air turbulence can lead to serious channel degradation [13]. Outdoor measurements just on channel deterioration have been conducted [11,16], but no confirmative conclusions could be obtained. That’s because of the challenges in measuring the impact of weather conditions over a long observation time and the difficulties in obtaining the re-occurrence of similar weather conditions to verify the results of independent measurements due to the time-varying rainfall rates and raindrop size distributions, as an example. Several groups have tried to build weather-emulating chambers in a controlled laboratory environment to generate reproducible weather for experimental measurements [13,17]. However, there are some distinctions between emulated and actual weather [8]. Therefore, a combined approach relying on both theoretical models and case studies is still valuable and necessary.
In this article, we present a comprehensive theoretical analysis that explores the biothermal heating effect of MMW/sub-THz waves on human skin under outdoor weather conditions. To the best of our knowledge, similar research has not yet been conducted, highlighting the importance of our investigation for future applications of this frequency band in outdoor settings. The methodology we propose can be adapted and applied to various weather conditions by measuring or determining the relevant parameters. Additionally, we emphasize that studying the thermal responses of rat skin provides a feasible approach to understanding the potential effects on human skin, considering their structural similarities. The article is organized as follows: Section 2 presents the dielectric properties of skin and water, and proposes theoretical models for the propagation of MMW/sub-THz waves in clear weather and rain. Section 3 shows theoretical models to predict the thermal heating effect and Section 4 presents theoretical models for thermal heating in outdoor weather with investigations on the influence of humidity and rainfall rate. Section 5 shows the conclusion and implications of this work.

2. Dielectric Properties and Channel Modeling

A description of skin response to MMW/sub-THz waves occurring outdoors requires detailed information about the skin and weather conditions (rain condition is considered here), such as the dielectric properties of skin and rainwater, temperature, humidity, and rainfall rate.

2.1. Dielectric Properties of Skin and Rainwater

The skin is the largest organ of mammals and its structure is composed of three interconnected layers of tissue: the epidermis as the outermost layer, the dermis as the middle layer, and the hypodermis as the innermost layer [18]. In this work, we consider the epidermis and dermis (with the thickness in the range of 0.06–0.1 mm and 1.2–2.8 mm, respectively [19]) in the theoretical model due to their ability to limit the penetration depth of MMW/sub-THz waves as discussed later. The dielectric properties of skin are characterized by measurements of its relative complex dielectric constant ε = ε j ε and have been proposed to be described by the single Debye model [20], the four Cole-Cole dispersion model [21] and other combined methods [22,23,24]. The four Cole-Cole model is considered the most widely used method based on open online databases [25]. It takes the skin to be a uniformly isotropic passive conducting medium [21] and can be expressed as
ε skin = ε + n = 1 4 Δ ε n 1 + ( j ω τ n ) 1 α n + σ j ω ε 0
Here, the parameter ω refers to the angular frequency and τ refers to the relaxation time. Parameter ε represents the dielectric constant in the high-frequency limit; α1, α2, α3, and α4 indicate the broadening due to dispersion; σ is the ionic conductivity of the skin; and ε0 represents the dielectric property of free space. For dry skin, the parameters are set as τ1 = 7.23 ps, τ2 = 32.48 ns, ε = 4.0, α1 = 0.0, α2 = 0.2, Δε1 = 32.0, Δε2 = 1100, and σ = 0.0002. For wet skin, parameters are τ1 = 7.96 ps, τ2 = 79.58 ns, τ3 = 1.59 μs, τ4 = 1.592 ms, ε = 4.0, α1 = 0.1, α2 = 0.0, α3 = 0.16, α4 = 0.2, Δε1 = 39.0, Δε2 = 280, Δε3 = 3.0 × 104, Δε4 = 3.0 × 104, and σ = 0.0004. The sweat duct inside the skin, with its radius ranging between 45 μm and 50 μm [26], is neglected due to its smaller size compared with the wavelength of MMW/sub-THz waves. We estimated the penetration depth by using the general expression p = (67.52/f[MHz]){[(ε′)2 + (ε″)2]1/2ε′}−1/2 based on Equation (1) and show its dependence on operating frequency in Figure 1a. The penetration depth decreases exponentially, which agrees well with the measured values set in the ICNIRP guidelines [2].
The dielectric properties of rainwater are mainly determined by the vibration mode of hydrogen bonds between water molecules. The bonds resonate and thus exert strong absorption on MMW/sub-THz waves. The absorbed energy is converted into kinetic energy for the irregular motion of water molecules and then transformed into thermal energy with an increase in mutual collision frequency [27]. The dielectric properties of water can be described by the Double-Debye dielectric model (D3M), which was first developed for saline water based on experimental measurements [28,29,30,31] and can also be reduced to pure water [32] in the frequency range up to 1000 GHz [33]. It can be expressed [34] as
ε water = ε w 0 f · n = 1 2 ε w ( n 1 ) ε wn f + j γ n
with ε w 0 ( T ) = 77.66 103.3 · T in as the static dielectric constant of water and T in = 1 300 / ( 273.15 + T ) as the modified relative inverse temperature variable. Parameters ε w 1 = 0.0671 ε w 0 and ε w 2 = 3.52 + 7.52 · T in are the first and second high-frequency constants (f→∞), respectively. Parameters γ 1 = 20.20 + 146.4 · T in + 316 · T in 2 and γ 2 = 39.8 γ 1 denote the primary and secondary relaxation frequencies. Tw is the water temperature in °C, and f is the operating frequency in GHz.
As shown in Figure 1a, the penetration depth in skin is always higher than that in water because of the smaller dielectric constant of skin [23]. The penetration depth is so sensitive to the incident frequency that higher frequencies cannot arrive at the skin surface if the skin has a thick water layer covering it due to falling rain droplets. A calculation of the transmissivity through the air-dry skin and water-wet skin interfaces is conducted and shown in Figure 1a by using Fresnel’s law with a horizontal monopolarization (hh) configuration. A higher transmission through the water-skin interface is observed, due to the smaller dielectric difference between air (free space) and skin on both sides as the water content of skin tissue can be up to 70% [35]. In the following sub-section, we conduct further calculations by assuming a layer of in vivo skin tissue instead of an in vitro one [36] and neglect the variation of its water content due to its self-adjusting capability.

2.2. Channel Models

In our analysis of rainy weather conditions, we consider the presence of rainwater adhering to bare skin, forming a uniform and planar water layer. When the power is focused onto a small area of the skin, both the skin and water layers can be treated as uniform and planar. The water layer introduces power losses, including reflection and absorption losses, on the MMW/sub-THz waves. We estimate the total power loss resulting from these effects by the equation
A w a t e r   ( dB ) = 10 log T t r
with Ttr being the total transmissivity through the water layer, which can be derived by using a multi-layer model [36]. As shown in Figure 1b, a three-layer structure (air-water-skin) is presented and can be estimated by the Fresnel principle simply, which has been confirmed in reference [36], when each layer is considered to be uniform. An expression for the total transmissivity from air into the skin can be expressed by the equation [37]
T t r = ( 1 + ρ 12 ) ( 1 + ρ 23 ) e j k 2 δ cos θ 2 1 + ρ 12 ρ 23 e j 2 k 2 δ cos θ 2 ,
where ρ12 and ρ23 are the reflection coefficients of the air-water interface and the water-skin interface, respectively. Parameter k2 is the wave number in the water layer, δ is the thickness of the water layer, and θ2 is the refraction angle in the water layer corresponding to the incident angle θ1, as shown in Figure 1b. The exponential term represents the propagation factor in the water layer between its top boundary (air-water interface) and bottom boundary (water-skin interface) along the angle θ2. Here, we neglect the scattering loss due to random wrinkles of the skin surface with a maximum depth of 0.3 mm [36], which is much smaller than the wavelength of the MMW/sub-THz waves. Thus, the total transmissivity relates to the reflection by the air-water interface and the water-skin interface, and also to the absorption by the water layer. Figure 1c shows a calculation of total transmissivity through the water layer with different incident angles. It increases first and then starts to decrease for frequencies above 100 GHz. By comparing it with the evolution curve of transmissivity in Figure 1a, we can attribute the difference between both trends to the absorption loss by the water layer. To further confirm this, we calculate the total transmissivity through a water layer with different thicknesses and plot it in Figure 1d. The total transmissivity increases with respect to frequency when the water layer is thinner than 0.2 mm, and follows an opposite trend when the thickness (δ) is larger than 0.2 mm, due to the higher absorption loss by the increased water layer thickness.
In clear weather, the gaseous composition of the atmosphere, mostly water vapor, strongly affects the propagation of MMW/sub-THz waves. The gaseous attenuation is commonly described as the summation of spectral absorption lines of water vapor and oxygen, and a water vapor continuum component. A line-by-line calculation is provided by the ITU recommendation sector (ITU-R) [38] based on the physical model MPM93 [39]. This method is proven to be valid at frequencies up to 500 GHz [40]. The results shown in Figure 2 indicate the ability to reduce the influence of atmospheric attenuation by setting the incident frequencies away from the water vapor absorption resonances. It should also be noted that higher terahertz frequencies are not preferred due to the large gaseous attenuation, especially when the relative humidity (RH) is higher than 60% in rainy weather.
The attenuation due to falling rain can generally be predicted by using the Mie scattering theory. This is reasonable because the ordinary size of raindrop particles is in the range from mm to cm [41], which is comparable to the MMW/sub-THz wavelength. The scattering characteristics can be described by the absorption and scattering coefficients. Two important underlying assumptions are usually considered: firstly, each scatter is independent of the other, and secondly, multiple scattering can be neglected. Under those assumptions, the attenuation suffered in rain [42] can be determined using the following equation
κ r a i n = 4 . 343 × 10 3 0 σ e x t ( n r , r , λ ) N ( r ) d r   ( dB / km )
where nr is the refractive index of the rain particles, r is the particle radius, λ is the carrier wavelength, and N(r) is the residual size distribution of raindrops, which is one of the main roles influencing the accuracy of theoretical predictions [8]. The parameter σ ext refers to the total extinction cross-section in the expression σ ext = σ abs + σ sca , with σabs and σsca as cross-sections of absorption and scattering, respectively. All three terms depend on the incident wavelength (λ), refractive index (nr), and radius (r) of water droplets. The power attenuation due to falling rain is calculated and shown in Figure 2. Here, we use the Marshall-Palmer [M-P] raindrop size distribution, which was also found to be reasonable for snow [43]. The Joss distribution was first obtained by measuring the size distribution of raindrops in Switzerland using a disdrometer [44]. The specific model proposed by ITU-R is based on measured data and empirical conduction [45]. A moderate falling rain (rain rate ranging from 2.6 to 7.5 mm/h) exerts a higher attenuation than the atmospheric gases when channels operate below 200 GHz. This indicates that the influence of falling rain on the biothermal heating effect cannot be neglected from the viewpoint of channel propagation. The Third Generation Partnership Project (3GPP) TS 38.104 has defined 24.25–52.6 GHz as frequency range 2 (FR2) for 5G NR [46]. In consideration of the significant potential demonstrated by certain terahertz frequency band windows, such as 140 GHz, 220 GHz, and 340 GHz, in various short-range wireless communication applications (e.g., V2X, high-speed transportation communication, and backhaul/fronthaul), we have included calculations for the 140 GHz frequency in addition to the 33.5 GHz frequency used in [47]. This selection is based on several factors, including lower transmission losses and higher penetration depth in water and skin layers, as illustrated in Figure 1 and Figure 2.

3. Thermal Heat Modeling

The biothermal heating effect by MMW/sub-THz waves on the skin is a complex process. The relevant parameters include frequency, power density, irradiation time, heat conduction, blood flow velocity in skin tissue, convection between skin surface and air, sweat evaporation and metabolic rate, etc. [48]. When electromagnetic powers are focused onto the skin surface, they will be absorbed and converted into heat, which is equivalent to establishing a heat source inside the skin tissue. The illumination spot inside the skin tissue acts as a heat source and the energy it generates can be expressed by the equation Q r ( x ) = ( I 0 Λ / L ) · e x / L [48], with x as the depth of the source, I0 as the illumination power density, and L as the penetration depth in the skin. The parameter Λ = | τ | 2 refers to the power transfer coefficient and can be calculated by Fresnel’s principle with the transmission coefficient as τ = 2η2cosθ1/(η2cosθ1 + η1cosθ2). Here parameters η1 and η2 are the wave impedance of electromagnetic waves in air and skin, respectively. Parameters θ1 refers to the angle of incidence and θ2 to the angle of reflection.
The classical Bioheat Transfer Equation (BHTE) models, such as Pennes’ BHTE, Weinbaum-Jiji (W-J) BHTE, and the thermal wave model of bioheat transfer (TWMBT), can be adapted to express the heat transfer mechanism inside the tissue. The Pennes’ BHTE model lays a solid and favorable foundation for establishing and studying the distribution of tissue temperature, due to its superiorities in simple structure, strong practicality, and wide range of applications. The W-J BHTE model is based on the fine anatomy of skin tissue [49] and takes into account the changes in number density, size, blood flow velocity, and direction of blood vessels in the tissue’s microstructure along with the penetration direction of electromagnetic waves. Its solution is more applicable to micro-vascular tissue due to its complexity but is not as efficient for a single skin layer. The TWMBT model considers the finite speed of heat transmission and can reflect the fluctuation characteristics of heat flow [50], which is not necessary or required here. In this work, we employ Pennes’ BHTE model, which is considered to be more reasonable in describing the microstructure and temperature distribution of subcutaneous areas caused by external heat sources [51]. It is also regarded as a common method to study the dependence of the biothermal heating effect on specific parameters, such as temperature, frequency, illumination time, and power density.
There are several assumptions for establishing the model: (1) the convective heat transfer process of blood occurs in the dermis with capillaries, (2) the tissue and blood in capillaries have the same temperature and are equal to the tissue temperature, and (3) the blood perfusion rate in local capillaries is isotropic. We also consider the skin tissue as a semi-infinite, homogeneous, and isotropic medium, with the heat exchange as a one-dimensional process, which is available to describe the heat transfer due to the thermal heating effect [52]. Thus, Pennes’ BHTE model can be simplified to be a one-dimensional form as
ρ c T t = k 2 T x 2 + W b c b ( T b T ) + Q m + Q r
where parameters ρ, c, and k are the density, specific heat, and thermal conductivity of the skin, respectively. Parameter Wb refers to the blood perfusion rate per unit volume with cb as the blood-specific heat and Tb as the arterial blood temperature. Parameter T is the skin temperature, Qm is metabolic heat production, and Qr is the electromagnetic wave heat source term. Under the assumption that the heat generated by the skin metabolism is negligible compared to the effect of blood perfusion on the skin temperature distribution [52], the term Qm in Equation (6) can be ignored. The corresponding initial and boundary conditions for Equation (6) are set as T ( x , t ) t = 0 = T 0 , T ( x , t ) x = L s = T c , and k T / x x = 0 = h ( T T e ) with T0, Tc, Te, Ls, and h representing the initial skin temperature, the core skin temperature, the ambient temperature, the skin thickness, and the total heat transfer coefficient between the skin tissue and the environment, respectively. The core skin temperature can also be called deep tissue temperature, which refers to the subcutaneous tissue temperature. Our work referred to the skin temperature range in [53,54], so 28.85 °C was selected as the initial skin temperature. Some of the values of these parameters used in the following calculations can be found in Table 1.
In order to check the efficiency of the model by Equation (6), we estimate the temporal evolution of temperature due to the illumination of a 33.5 GHz channel onto a layer of skin, with the power density changing from 284 mW/cm2 to 664 mW/cm2. In Figure 3a, the variation of skin temperature versus illumination time is plotted. The temperature increases significantly for longer illumination time and higher power density and then comes to be a constant (equilibrium). The calculation agrees well with the measured data obtained by Wang [47], who employed a layer of in vitro rat skin in the laboratory under the illumination of a continuous wave at 33.5 GHz and launched out from a lens antenna. The agreement between our calculations and the measured data indicates the efficacy of Pennes’ BHTE model in analyzing the temperature evolution of rat skin under electromagnetic radiation, provided that the total heat transfer coefficient (h) is determined as shown in Figure 3b. These results support the transferability of our findings to discussions on the thermal heating effect of human skin, despite its thickness being 5–6 times greater than that of rat skin. It is important to consider that as body temperature rises, mammals regulate their temperature through physiological activities, including increased heat dissipation, which results in an augmented heat transfer between the skin and the surrounding environment. Consequently, the heat transfer coefficient increases in response. Based on the curves shown in Figure 3b, which are calculated by fitting the measured data in Figure 3a, we can see that the coefficient h evolves with ∆T along the same trend before stabilizing (hmax). This means there should be a correlation between the heat transfer coefficient and the temperature difference ∆T (between the skin and ambient temperature), which has been demonstrated theoretically in reference [57].
To see the influence of a rainwater layer on human skin on a rainy day, we predicted the variation of thermal heating with/without a planar water layer by employing Equation (6). In the case of water layer coverage, we adopted the dielectric properties of wet skin. The heat transfer coefficient of rats in Figure 3b was used here because it has been proven to be close to that of human skin [58]. Calculations were completed using the parameters in Table 1 and setting the thickness of the planar water layer to 0.2 mm due to the relatively stable transmissivity over the whole frequency range as indicated in Figure 1d. The results are shown in Figure 3c. It is obvious that the temperature increases faster when there is a water layer. We attribute this to the insulation effect of the water layer, which has a much higher specific heat capacity than air [59]. It should also be noted that the temperature increases with respect to the operating frequency. This is due to the larger absorption coefficient by the uniform and planar skin layer at higher frequencies, as indicated in Figure 6(b) of reference [36]. The same phenomenon can also be observed in Figure 3d, which shows a temporal variation of human skin temperature due to the illumination of channels operating at 33.5 GHz and 140 GHz. However, this does not mean a higher frequency is preferred for thermal heating devices because it suffers limitations of low penetration depth, as shown in Figure 1a; high propagation loss, as shown in Figure 2; and low output power [60].

4. Outdoor Thermal Heating Performance

To evaluate the thermal heating performance of MMW/sub-THz waves in rainy weather, we combined the theoretical models for thermal heating by Equation (6), channel propagation by Equations (3)–(5), and free space path loss (FSPL). Calculations were conducted by setting a rain rate of 5 mm/hr and a planar and uniform water layer of 0.2 mm covering the surface of human skin surface due to falling rain, a normal incidence configuration for the largest transmissivity, as shown in Figure 1c, as well as a relative humidity (RH) of 30% in clear weather and 60% in rain, output power of 60 dBm at the transmitter side launched out from an antenna with a gain of 50 dBi, and an illumination time of 240 s. Results are plotted in Figure 4.
A negative correlation between skin temperature and channel propagation distance is shown in Figure 4a, due to the distance dependence of channel propagation losses, as shown in Figure 2. The temperature increase caused by the 140 GHz channel is smaller than that by the 33.5 GHz channel due to the higher propagation losses and lower transmissivity into the skin tissue, as shown in Figure 1a. This means the thermal heating effect by sub-THz frequencies would not be an urgent concern for human skin if the propagation distance or output power density is limited. It should also be noted that the equilibrium temperature in rain is much lower than that in clear weather, which is opposite to our predictions in Figure 3d, due to the power losses caused by the water layer covering the skin surface. This indicates that the influence of falling rain and water vapor humidity is not so obvious when compared with power loss caused by the water layer.
To further investigate the dependence of the equilibrium temperature by the 33.5 GHz signal on the rainfall rate and relative humidity, we calculated evolutional curves due to the variation of rainfall rate and humidity and show the results in Figure 4b,c respectively. The horizontal lines imply the negligible influence of rain conditions, which exert negligible propagation degradation compared to the FSPL and the power losses due to the presence of a water layer on the skin surface. This means the major deterioration of the thermal-heating effect in rain is from the rainwater falling on the skin surface and the output power density at the transmitter side.

5. Conclusions

The evaluation of the biothermal heating effect on human skin is crucial for the application of MMW/sub-THz waves in future wireless communications and biological weapon scenarios in outdoor environments. In this study, we extensively investigate the efficiency of MMW/sub-THz waves for biothermal heating under outdoor weather conditions, considering the challenges posed by water vapor and falling rain. To accomplish this, we propose a comprehensive theoretical model that incorporates Pennes’ bioheat transfer equation (BHTE), the Mie scattering theory, and the ITU-R model. The proposed model is validated using measured data and is found to accurately evaluate the heating effect. We systematically explore the impact of various factors, including water vapor humidity, rainfall rate, propagation distance, operating frequency, illumination time, and power density, on the biothermal heating effect. We observed that the heat transfer coefficient is highly sensitive to the temperature differences between the skin and the surrounding environment. In the presence of falling rain, the thermal heating effect of MMW/sub-THz waves is diminished due to the formation of a water layer on the skin surface. However, this water layer does not cause significant channel propagation losses when the distance between the power source and the target (skin) is limited.
It is important to note that our work has been validated by measurements conducted on rat skin, and the propagation models employed in our analysis have been widely justified. Although there are certain assumptions in our study with limited justifications, the persuasive nature of this work lies in the validation of the thermal heating model using rat skin measurements and the extensive justification of the propagation models. Consequently, we are confident in the reliability and validity of our findings.
In conclusion, this work confirms the current and near-future applications of sub-THz frequencies in biothermal heating. The developed theoretical model, validated by rat skin measurements, is also deemed suitable for evaluating in vitro-in vivo terahertz communication channels.

Author Contributions

Conceptualization, J.M.; methodology, M.W., P.L. and J.M.; software, M.W. and P.L.; validation, M.W.; investigation, M.W., P.L. and Y.L.; writing—original draft preparation, M.W.; writing—review and editing, J.M. and X.B.; supervision, J.M.; project administration, J.M.; funding acquisition, J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported in part by the National Natural Science Foundation of China (62071046), the Graduate Innovative Practice Project of Tangshan Research Institute, BIT (TSDZXX202201), the Science and Technology Innovation Program of Beijing Institute of Technology (2022CX01023) and the Talent Support Program of Beijing Institute of Technology “Special Young Scholars” (3050011182153).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We appreciate the support from Houjun Sun from the Beijing Institute of Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Popov, V.; Shevchenko, A. Analysis of standards and norms of electromagnetic irradiation levels in wireless communication systems on railway transport. Procedia Comput. Sci. 2019, 149, 239–245. [Google Scholar] [CrossRef]
  2. International Commission on Non-Ionizing Radiation Protection (ICNIRP). Guidelines for limiting exposure to electromagnetic fields (100 kHz to 300 GHz). Health Phys. 2020, 118, 483–524. [Google Scholar] [CrossRef] [PubMed]
  3. Hirata, A.; Diao, Y.; Onishi, T.; Sasaki, K.; Ahn, S.; Colombi, D.; Santis, V.D.; Laakso, I.; Giaccone, L.; Joseph, W.; et al. Assessment of human exposure to electromagnetic fields: Review and future directions. IEEE Trans. Electromagn. Compat. 2021, 63, 1619–1630. [Google Scholar] [CrossRef]
  4. Bailey, W.H.; Bodemann, R.; Bushberg, J.; Chou, C.; Cleveland, R.; Faraone, A.; Foster, K.R.; Gettman, K.E.; Graf, K.; Harrington, T.; et al. Synopsis of IEEE Std C95. 1™-2019 “IEEE Standard for Safety Levels with Respect to Human Exposure to Electric, Magnetic, and Electromagnetic Fields, 0 Hz to 300 GHz”. IEEE Access 2019, 7, 171346–171356. [Google Scholar] [CrossRef]
  5. Sasaki, K.; Mizuno, M.; Wake, K.; Watanabe, S. Monte Carlo simulations of skin exposure to electromagnetic field from 10 GHz to 1 THz. Phys. Med. Biol. 2017, 62, 6993–7010. [Google Scholar] [CrossRef]
  6. Kanezaki, A.; Hirata, A.; Watanabe, S.; Shirai, H. Effects of dielectric permittivities on skin heating due to millimeter wave exposure. Biomed. Eng. Online 2009, 8, 20. [Google Scholar] [CrossRef] [Green Version]
  7. Hashimoto, Y.; Hirata, A.; Morimoto, R.; Aonuma, S.; Laakso, I.; Jokela, K.; Foster, K.R. On the averaging area for incident power density for human exposure limits at frequencies over 6 GHz. Phys. Med. Biol. 2017, 62, 3124. [Google Scholar] [CrossRef]
  8. Li, P.; Wang, J.; Zhao, L.; Ma, J.; Sun, H.; Moeller, L.; Federici, J.F. Performance degradation of terahertz channels in emulated rain. Nano Commun. Netw. 2023, 35, 100431. [Google Scholar] [CrossRef]
  9. Wang, R.; Mei, Y.; Meng, X.; Ma, J. Secrecy performance of terahertz wireless links in rain and snow. Nano Commun. Netw. 2021, 28, 100350. [Google Scholar] [CrossRef]
  10. Amarasinghe, Y.; Zhang, W.; Zhang, R.; Mittleman, D.M.; Ma, J. Scattering of Terahertz Waves by Snow. J. Infrared Millim. Terahertz Waves 2020, 41, 215–224. [Google Scholar] [CrossRef]
  11. Ma, J.; Adelberg, J.; Shrestha, R.; Moeller, L.; Mittleman, D.M. The Effect of Snow on a Terahertz Wireless Data Link. J. Infrared Millim. Terahertz Waves 2018, 39, 505–508. [Google Scholar] [CrossRef]
  12. Ma, J.; Shrestha, R.; Moeller, L.; Mittleman, D.M. Invited Article: Channel performance for indoor and outdoor terahertz wireless links. APL Phontonics 2018, 3, 51601. [Google Scholar] [CrossRef] [Green Version]
  13. Federici, J.F.; Ma, J.; Moeller, L. Review of weather impact on outdoor terahertz wireless communication links. Nano Commun. Netw. 2016, 10, 13–26. [Google Scholar] [CrossRef] [Green Version]
  14. Ma, J.; Vorrius, F.; Lamb, L.; Moeller, L.; Federici, J.F. Experimental Comparison of Terahertz and Infrared Signaling in Laboratory-Controlled Rain. J. Infrared Millim. Terahertz Waves 2015, 36, 856–865. [Google Scholar] [CrossRef]
  15. Ma, J.; Vorrius, F.; Lamb, L.; Moeller, L.; Federici, J.F. Comparison of Experimental and Theoretical Determined Terahertz Attenuation in Controlled Rain. J. Infrared Millim. Terahertz Waves 2015, 36, 1195–1202. [Google Scholar] [CrossRef]
  16. Hirata, A.; Yamaguchi, R.; Takahashi, H.; Kosugi, T.; Murata, K.; Kukutsu, N.; Kado, Y. Effect of Rain Attenuation for a 10-Gb/s 120-GHz-Band Millimeter-Wave Wireless Link. IEEE Trans. Microw. Theory Technol. 2009, 57, 3099–3105. [Google Scholar] [CrossRef]
  17. Yang, Y.; Mandehgar, M.; Grischkowsky, D.R. Broadband THz Signals Propagate Through Dense Fog. IEEE Photonics Technol. Lett. 2015, 27, 383–386. [Google Scholar] [CrossRef]
  18. Nikitkina, A.I.; Bikmulina, P.Y.; Gafarova, E.R.; Kosheleva, N.V.; Efremov, Y.M.; Bezrukov, E.A.; Butnaru, D.V.; Dolganova, I.N.; Chernomyrdin, N.V.; Cherkasova, O.P.; et al. Terahertz radiation and the skin: A review. J. Biomed. Opt. 2021, 26, 43005. [Google Scholar] [CrossRef]
  19. Wu, T.; Rappaport, T.S.; Collins, C.M. Safe for Generations to Come: Considerations of Safety for Millimeter Waves in Wireless Communications. IEEE Microw. Mag. 2015, 16, 65–84. [Google Scholar] [CrossRef] [Green Version]
  20. Gandhi, O.P.; Riazi, A. Absorption of Millimeter Waves by Human Beings and its Biological Implications. IEEE Trans. Microw. Theory Technol. 1986, 34, 228–235. [Google Scholar] [CrossRef] [Green Version]
  21. Gabriel, S.; Lau, R.W.; Gabriel, C. The dielectric properties of biological tissues: III. Parametric models for the dielectric spectrum of tissues. Phys. Med. Biol. 1996, 41, 2271–2293. [Google Scholar] [CrossRef] [Green Version]
  22. Pickwell, E.; Cole, B.E.; Fitzgerald, A.J.; Wallace, V.P.; Pepper, M. Simulation of terahertz pulse propagation in biological systems. Appl. Phys. Lett. 2004, 84, 2190–2192. [Google Scholar] [CrossRef]
  23. Alekseev, S.I.; Ziskin, M.C. Human skin permittivity determined by millimeter wave reflection measurements. Bioelectromagnetics 2007, 28, 331–339. [Google Scholar] [CrossRef]
  24. Chahat, N.; Zhadobov, M.; Augustine, R.; Sauleau, R. Human skin permittivity models for millimetre-wave range. Electron. Lett. 2011, 47, 427–428. [Google Scholar] [CrossRef]
  25. Hasgall, P.A.; Gennaro, F.D.; Baumgartner, C.; Neufeld, E.; Lloyd, B.; Gosselin, M.C.; Klingenbock, A.; Kuster, N. IT’ IS Database for Thermal and Electromagnetic Parameters of Biological Tissues. Available online: https://itis.swiss/virtual-population/tissue-properties/overview/ (accessed on 10 July 2023).
  26. Tripathi, S.R.; Miyata, E.; Ishai, P.B.; Kawase, K. Morphology of human sweat ducts observed by optical coherence tomography and their frequency of resonance in the terahertz frequency region. Sci. Rep. 2015, 5, 9071. [Google Scholar] [CrossRef] [Green Version]
  27. Kristensen, T.T.; Withayachumnankul, W.; Jepsen, P.U.; Abbott, D. Modeling terahertz heating effects on water. Opt. Express 2010, 18, 4727–4739. [Google Scholar] [CrossRef] [Green Version]
  28. Barthel, J.; Bachhuber, K.; Buchner, R.; Hetzenauer, H. Dielectric spectra of some common solvents in the microwave region. Water and lower alcohols. Chem. Phys. Lett. 1990, 167, 62–66. [Google Scholar] [CrossRef]
  29. Kindt, J.T.; Schmuttenmaer, C.A. Far-Infrared Dielectric Properties of Polar Liquids Probed by Femtosecond Terahertz Pulse Spectroscopy. J. Phys. Chem. 1996, 100, 10373–10379. [Google Scholar] [CrossRef]
  30. Stogryn, A.P.; Bull, H.T.; Rubayi, K.; Iravanchy, S. The microwave permittivity of sea and freshwater. In Arrojet Internal Report; GenCorp Aerojet: Sacramento, CA, USA, 1996. [Google Scholar]
  31. Ronne, C.; Thrane, L.; Astrand, P.O.; Wallqvist, A.; Mikkelsen, K.V.; Keiding, S.R. Investigation of the temprerature dependence of dielectric relaxation in liquid water by THz reflection spectroscopy and molecular dynamics simulation. J. Chem. Phys. 1997, 107, 5319–5331. [Google Scholar] [CrossRef]
  32. Mätzler, C. Thermal Microwave Radiation: Applications for Remote Sensing; Institution of Engineering & Technology: Stevenage, UK, 2007. [Google Scholar]
  33. Ulaby, F.T.; Moore, R.K.; Fung, A.K. Microwave Remote Sensing: Actice and Passive. In Radar Remote Sensing and Surface Scattering and Emission Theory; Artech House: Norwood, MA, USA, 1982; Volume 2. [Google Scholar]
  34. Liebe, H.J.; Hufford, G.A.; Manabe, T. A model for the complex permittivity of water at frequencies below 1 THz. Int. J. Infrared Millim. Waves 1991, 12, 659–675. [Google Scholar] [CrossRef]
  35. Nakagawa, N.; Matsumoto, M.; Sakai, S. In vivo measurement of the water content in the dermis by confocal Raman spectroscopy. Skin Res. Technol. 2010, 16, 137–141. [Google Scholar] [CrossRef] [PubMed]
  36. You, B.; Chen, C.; Yu, C.; Wang, P.; Lu, J. Frequency-dependent skin penetration depth of terahertz radiation determined by water sorption–desorption. Opt. Express 2018, 26, 22709–22721. [Google Scholar] [CrossRef] [PubMed]
  37. Ulaby, F.T.; Moore, R.K.; Fung, A.K. Microwave Remote Sensing: Actice and Passive. In Fundamentals and Radiometry; Artech House: Norwood, MA, USA, 1981; Volume 1. [Google Scholar]
  38. ITU-R Recommendation P.676-13: Attenuation by Atmospheric Gases and Related Effects. Available online: https://www.itu.int/rec/R-REC-P.676-13-202208-I/en (accessed on 10 July 2023).
  39. Liebe, H.J.; Hufford, G.A.; Cotton, M.G. Propagation modeling of moist air and suspended water/ice particles at frequencies below 1000 GHz. In Proceedings of the AGARD 52th Specialists’ Meeting of the Electromagnetic Wave Propagation Panel, Palma de Mallorca, Spain, 17–20 May 1993. [Google Scholar]
  40. O’Hara, J.F.; Grischkowsky, D.R. Comment on the Veracity of the ITU-R Recommendation for Atmospheric Attenuation at Terahertz Frequencies. IEEE Trans. Terahertz Sci. Technol. 2018, 8, 372–375. [Google Scholar] [CrossRef]
  41. Pruppacher, H.R.; Klett, J.D. Microphysics of Clouds and Precipitation; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1997. [Google Scholar]
  42. Deirmendjian, D. Electromagnetic Scattering on Spherical Polydispersions; American Elsevier Publishing: New York, NY, USA, 1969. [Google Scholar]
  43. Marshall, J.S.; Palmer, W.M. The distribution of raindrops with size. J. Meteorol. 1948, 5, 165–166. [Google Scholar] [CrossRef]
  44. Joss, J.; Thams, J.C.; Waldvogel, A. The variation of rain drop size distribution at Locarno. Bull. Am. Meteorol. Soc. 1968, 49, 2. [Google Scholar]
  45. ITU-R Recommendation P.838-3: Specific Attenuation Model for Rain for Use in Prediction Methods. 2005. Available online: https://www.itu.int/rec/R-REC-P.838/en (accessed on 10 July 2023).
  46. 3GPP TS 38.104. Base Station (BS) Radio Transmission and Reception. Available online: https://www.3gpp.org/ftp/Specs/archive/38_series/38.104 (accessed on 10 July 2023).
  47. Wang, W.Y. Study on Animal Surface Temperature Irradiated by Millimeter Wave; University of Electronic Science and Technology of China: Chengdu, China, 2005. [Google Scholar]
  48. Foster, K.R.; Kritikos, H.N.; Schwan, H.P. Effect of Surface Cooling and Blood Flow on the Microwave Heating of Tissue. IEEE Trans. Biomed. Eng. 1978, 25, 313–316. [Google Scholar] [CrossRef]
  49. Song, W.J.; Weinbaum, S.; Jiji, L.M. A Theoretical Model for Peripheral Tissue Heat Transfer Using the Bioheat Equation of Weinbaum and Jiji. J. Biomech. Eng. 1987, 109, 72–78. [Google Scholar] [CrossRef]
  50. Liu, J.; Ren, Z.P.; Wang, C.C. Interpretation of living tissue’s temperature oscillations by thermal wave theory. Chin. Sci. Bull. 1995, 12, 1493–1495. [Google Scholar]
  51. Pennes, H.H. Analysis of Tissue and Arterial Blood Temperatures in the Resting Human Forearm. J. Appl. Physiol. 1984, 85, 5–34. [Google Scholar] [CrossRef]
  52. Wang, H.; Burgei, W.A.; Zhou, H. Analytical solution of one-dimensional Pennes’ bioheat equation. Open Phys. 2020, 18, 1084–1092. [Google Scholar] [CrossRef]
  53. Musat, R.; Helerea, E. Parameters and Models of the Vehicle Thermal Comfort. Acta Univ. Sapientiae 2009, 1, 215–226. [Google Scholar]
  54. Benedict, F.G.; Miles, W.R.; Johnson, A. The Temperature of the Human Skin. Proc. Natl. Acad. Sci. USA 1919, 5, 218–222. [Google Scholar] [CrossRef]
  55. Torvi, D.A.; Dale, J.D. A Finite Element Model of Skin Subjected to a Flash Fire. J. Biomech. Eng. 1994, 116, 250–255. [Google Scholar] [CrossRef]
  56. Deng, Z.S.; Liu, J. Parametric studies on the phase shift method to measure the blood perfusion of biological bodies. Med. Eng. Phys. 2000, 22, 693–702. [Google Scholar] [CrossRef]
  57. Ishigaki, H.; Horikoshi, T.; Uematsu, T.; Sahashi, M.; Tsuchikawa, T.; Mochida, T.; Hieda, T.; Isoda, N.; Kubo, H. Experimental study on convective heat transfer coefficient of the human body. J. Therm. Biol. 1993, 18, 455–458. [Google Scholar] [CrossRef]
  58. Gilhar, A.; Etzioni, A. The nude mouse model for the study of human skin disorders. Dermatology 1994, 189, 5–8. [Google Scholar] [CrossRef]
  59. Tolgyessy, J. Chemistry and Biology of Water, Air and Soil: Environmental Aspects; Elsevier: Amsterdam, The Netherlands, 1993. [Google Scholar]
  60. Asada, M.; Suzuki, S. Terahertz emitter using resonant-tunneling diode and applications. Sensors 2021, 21, 1384. [Google Scholar] [CrossRef]
Figure 1. (a) Penetration depth in dry skin, wet skin, and water layers; transmissivity through the air-dry skin and water-wet skin interfaces with normal incidence. (b) Basic geometry of transmission through a water layer and into the skin. (c) Transmissivity through a water layer covered on the skin surface under different incident angles. (d) Transmissivity through a water layer (with a thinness of 0.2 mm) on the skin under different water layer thicknesses with an incident angle of 30°.
Figure 1. (a) Penetration depth in dry skin, wet skin, and water layers; transmissivity through the air-dry skin and water-wet skin interfaces with normal incidence. (b) Basic geometry of transmission through a water layer and into the skin. (c) Transmissivity through a water layer covered on the skin surface under different incident angles. (d) Transmissivity through a water layer (with a thinness of 0.2 mm) on the skin under different water layer thicknesses with an incident angle of 30°.
Applsci 13 08305 g001
Figure 2. Atmospheric gaseous attenuation under different values of relative humidity (RH) (left side) and attenuation due to rain under a rainfall rate of 5 mm/h when M-P, Joss, and ITU-R models are employed (right side).
Figure 2. Atmospheric gaseous attenuation under different values of relative humidity (RH) (left side) and attenuation due to rain under a rainfall rate of 5 mm/h when M-P, Joss, and ITU-R models are employed (right side).
Applsci 13 08305 g002
Figure 3. (a) Temperature evolution with respect to illumination time (Hollow circles for experimental data and solid curves for fitted values based on Equation (6)). (b) Response of the heat transfer coefficient on the variation of the temperature difference between the skin and environment. The colors correspond to the same power density as in (a). (c) Calculated evolution of temperature due to electromagnetic illumination with respect to frequency. The colors correspond to the same power density as in (a). (d) Calculated temperature evolution with respect to illumination time under a power density of 284 mW/cm2 at 33.5 GHz (blue) and 140 GHz (red). The solid and dashed curves correspond to the same weather conditions as in (c).
Figure 3. (a) Temperature evolution with respect to illumination time (Hollow circles for experimental data and solid curves for fitted values based on Equation (6)). (b) Response of the heat transfer coefficient on the variation of the temperature difference between the skin and environment. The colors correspond to the same power density as in (a). (c) Calculated evolution of temperature due to electromagnetic illumination with respect to frequency. The colors correspond to the same power density as in (a). (d) Calculated temperature evolution with respect to illumination time under a power density of 284 mW/cm2 at 33.5 GHz (blue) and 140 GHz (red). The solid and dashed curves correspond to the same weather conditions as in (c).
Applsci 13 08305 g003
Figure 4. (a) Temperature evolution versus channel distance (between transmitter and skin surface) due to the radiation of a 60 dBm output power from a transmitter over an illumination time of 240 s. (b) Temperature variation with respect to rainfall rate under different radiation power densities (at 33.5 GHz) over a propagation distance of 10 m. (c) Temperature variation versus relative humidity under different water layer thicknesses.
Figure 4. (a) Temperature evolution versus channel distance (between transmitter and skin surface) due to the radiation of a 60 dBm output power from a transmitter over an illumination time of 240 s. (b) Temperature variation with respect to rainfall rate under different radiation power densities (at 33.5 GHz) over a propagation distance of 10 m. (c) Temperature variation versus relative humidity under different water layer thicknesses.
Applsci 13 08305 g004aApplsci 13 08305 g004b
Table 1. Values of Physical Parameters for Theoretical Calculation [25,51,53,54,55,56].
Table 1. Values of Physical Parameters for Theoretical Calculation [25,51,53,54,55,56].
ParameterValueParameterValue
ρ (kg/m3)1200T0 (°C)28.85
c (J/kg/°C)3580Tc (°C)36.5
k (W/m/°C)0.4Te (°C)12.5
Wb (kg/s/m3)0.59Ls (cm)1.0
cb (J/kg/°C)4200h (W/m2·K)13
Tb (°C)36.5
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

Wei, M.; Li, P.; Lei, Y.; Bao, X.; Ma, J. Biothermal Heating on Human Skin by Millimeter and Sub-Terahertz Waves in Outdoor Environment—A Theoretical Study. Appl. Sci. 2023, 13, 8305. https://doi.org/10.3390/app13148305

AMA Style

Wei M, Li P, Lei Y, Bao X, Ma J. Biothermal Heating on Human Skin by Millimeter and Sub-Terahertz Waves in Outdoor Environment—A Theoretical Study. Applied Sciences. 2023; 13(14):8305. https://doi.org/10.3390/app13148305

Chicago/Turabian Style

Wei, Menghan, Peian Li, Yuanshuai Lei, Xiue Bao, and Jianjun Ma. 2023. "Biothermal Heating on Human Skin by Millimeter and Sub-Terahertz Waves in Outdoor Environment—A Theoretical Study" Applied Sciences 13, no. 14: 8305. https://doi.org/10.3390/app13148305

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