Next Article in Journal / Special Issue
Reply to Taylor et al. Comment on “Manna et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443”
Previous Article in Journal
Effect of a Second Pregnancy on the HPV Serology in Mothers Followed Up in the Finnish Family HPV Study
 
 
Reply published on 18 October 2023, see Viruses 2023, 15(10), 2111.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Comment

Comment on Manna et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443

by
Gavin J. Taylor
1,*,
Jérémie Margueritat
2 and
Lucien Saviot
3
1
Institute for Globally Distributed Open Research and Education (IGDORE), São Carlos 13560-230, Brazil
2
Institut Lumière Matière, UMR5306, Université de Lyon, 69622 Villeurbanne, France
3
Laboratoire Interdisciplinaire Carnot de Bourgogne, UMR 6303 CNRS–Université de Bourgogne Franche Comté, 21000 Dijon, France
*
Author to whom correspondence should be addressed.
Viruses 2023, 15(10), 2110; https://doi.org/10.3390/v15102110
Submission received: 9 August 2023 / Revised: 31 August 2023 / Accepted: 7 September 2023 / Published: 18 October 2023
In a recent article published in Viruses by Manna et al. [1], the authors estimate the resonant frequency of SARS-CoV-2 particles using elastic continuum theory of the vibrational modes of a sphere [2]. Unfortunately, the article provides an incorrect solution to the eigenvalue equation. Although this does not influence the validity of the experimental results, the error will be misleading if the solution is used to select frequencies for testing virus inactivation in other studies. This comment provides details on how to obtain a correct solution numerically.
Elastic continuum theory can be used to calculate the resonant frequencies of the vibrational modes of a homogenous isotropic free sphere [2]. This is often simply referred to as Lamb’s theory or equation, and only requires the radius of the sphere, as well as its longitudinal and transverse sound velocities, as input (Appendix A). Recent studies have investigated microwave resonant absorption as a mechanism for inactivating spherical viruses [3,4,5,6,7,8]. These studies propose that virus inactivation results from a structural-resonant energy transfer (SRET) process, where electromagnetic fields matching the frequency of the dipolar mode of a virus capsid excite a confined acoustic vibration which damages its structure [9]. Although most inactivation studies have empirically tested the effectiveness of a range of microwave frequencies, the frequency of the dipolar mode can be measured directly using microwave absorption spectroscopy [9,10,11,12,13,14]. The peak absorption frequency from spectroscopy can then be used as the centre of a frequency range to test for inactivation experiments [7,9,10]. Alternatively, Lamb’s equation can be used to calculate the dipolar resonant frequency across the size range of the virus, which indicates a frequency range over which inactivation can be tested [1,15]. However, while the dimensions of many viruses have been measured using electron microscopy, acoustic properties are only available for a single virus (STMV [16]). The applicability of Lamb’s equation to spherical viruses is also limited, as they are inhomogeneous, many are only approximately spherical, and experiments are usually performed with viruses suspended in a viscous liquid [17]. Nonetheless, Lamb’s theory can still provide a useful estimate of the dipolar resonance frequency for designing or analysing inactivation experiments.
To estimate the resonant frequency (f) of SARS-CoV-2, Manna et al. [1] derived an analytical solution to Lamb’s eigenvalue equation for the first harmonic of the dipolar mode of a sphere (their Equation (1)). Their solution starts by assuming that the ratio between the longitudinal and transverse sound velocities is 2, and uses an approximate expression for the asymptotic value of a Bessel function when its argument approaches zero (their Equation (2)). Manna et al. [1] used this approximation to replace the spherical Bessel functions in the eigenvalue equation with polynomial functions of the eigenvalues ξ and η (for completeness, we note that the approximation for small arguments of a spherical Bessel functions is: j m z z m 2 m + 1 [18] (p. 708), while their work uses the approximation for cylindrical Bessel functions). However, after performing this substitution, they omitted the multiplication by ξ (which is present in their Equation (1)) from the first term of their Equation (3). Consequently, their Equation (4) should be corrected to η2 = 2ξ2. Substituting η = 2ξ into the corrected Equation (4) then leads to ξ = 0, which produces a result of f = 0 regardless of the radius or sound velocity. The same error has also been made elsewhere [15]. Although trivial, the 0 Hz solution is, in fact, correct, and corresponds to translations of the sphere. However, this approach does not indicate the resonant frequency of the dipolar mode, which is the desired result for selecting the frequency range to test in microwave inactivation experiments. A better approach is to solve the eigenvalue equation numerically (Appendix A).
We provide a python script which numerically calculates the solution for the frequencies of the vibrational modes of a sphere (github.com/gavinscode/sphere_resonance [19] (accessed on 5 August 2023)). Additionally, a web app (saviot.cnrs.fr/lamb, [20] (accessed on 5 August 2023)) can be used to calculate the frequencies and visualise the corresponding displacements of the resonant modes. Both of these tools can be used regardless of the ratio between longitudinal and transverse sound velocities, and they can also calculate the frequencies of all the spheroidal resonant modes and harmonics (Appendix A).
Following Manna et al. [1], we calculated the resonant frequencies of virus particles from 30 to 70 nm in radius (60 to 140 nm in diameter [21]), with longitudinal and transverse sound velocities of 1800 m/s and 900 m/s, and estimated that the resonant frequency of SARS-CoV-2 is between 7.4 to 17.2 GHz. Using Equation (5) from Manna et al. [1] with the parameters above provided a frequency range of 12.9 to 30 GHz, although the initial estimate of 6.5 to 16 GHz seems to be based on mistakenly using the virus diameter in place of its radius in Equation (5). In their preliminary experiments, Manna et al. [1] tested inactivation of SARS-CoV-2 with frequencies from 6.5 to 17 GHz, and found that inactivation peaked at 10 GHz. The SRET paradigm suggests that this inactivation peak corresponds to the dipolar resonant frequency of the virus, although thermal mechanisms may have contributed to the observed inactivation, and additional analysis is required in order to exclude these [3]. However, Manna et al. [1] would not have found this inactivation peak if they had only tested frequencies across the range of 12.9 to 30 GHz resulting from their Equation (5), whereas our solution does indicate a range (7.4 to 17.2 GHz) that includes the 10 GHz peak. As such, we recommend that future research on virus inactivation use a numerical solution for Lamb’s equation when an estimate of the dipolar resonant frequency is required.

Funding

L.S. acknowledges support from the EIPHI Graduate School (Contract ANR-17-EURE-0002).

Conflicts of Interest

G.T. provides consulting services for Resonant Health Inc.

Appendix A

Lamb’s theory [2] describes the free vibrations of an elastic sphere with stress-free boundary conditions. Two types of vibrational modes exist: torsional (TOR) modes are purely transverse displacement, while spherical (SPH) modes involve radial displacements. Only the latter will be covered here. Vibrational modes are further described by an angular momentum index (l) and a harmonic index (n), and the terminology (SPH, l, n) is often used to indicate a spherical mode with a specific angular momentum and harmonic. For example, (SPH, l = 1, n = 2), or simply (SPH, 1, 2), indicates the spherical mode with an angular momentum index of one and its second harmonic (note that some authors start indexing harmonics from zero, while others start at one).
The frequencies of the vibrational modes can be found by solving the eigenvalue equation [2]:
4 η 2 j l η + ( l 1 ) ( l + 2 ) j l + 1 η η ( l + 1 ) j l η j l + 1 ξ ξ + η 4 + 2 ( l 1 ) ( 2 l + 1 ) η 2 j l η + 2 η 2 2 l ( l 1 ) ( l + 2 ) j l + 1 η η j l ξ = 0
when l ≥ 1, or
4 V t 2 j 1 ξ V l 2 ξ j 0 ξ = 0
when l = 0.
The eigenvalues are ξ = 2 π R f V l and η = 2 π R f V t , where R is the radius of the sphere, Vl and Vt are the longitudinal and transverse sound velocities of the sphere, and f is the frequency. jl are spherical Bessel functions of the first kind, where the order of the Bessel function is referenced using the angular momentum index (i.e., l or l + 1). Note that we have multiplied the usual expression for the eigenvalue equation (e.g., Equation (1) in the supplementary material of Yang et al. [9]) by jl(η)jl(ξ) so that no terms are divided by a Bessel function. This removes discontinuities from the results and avoids numerical issues with dividing by zero, without changing the frequencies of the roots. The Bessel functions required to solve for l = 0, 1, and 2 are [18] (pp. 701–702):
j 0 z = s i n z   z j 1 z = s i n z   z 2     c o s z   z j 2 z = 3 z 2 1 s i n z   z     3 c o s z   z 2 j 3 z = 15 z 3 6 z s i n z   z     15 z 2 1 c o s z   z
The l = 0 modes from solving Equation (A2) are termed breathing modes. When l = 1, the solutions are termed dipolar modes, and Equation (A1) can be simplified to:
4 j 2 ξ ξ j 1 η + η 2 j 1 η + 2 j 2 η η j 1 ξ = 0
When l = 2, the solutions are termed quadrupolar modes, and Equation (A1) can be simplified to:
( 4 η 2 48 ) j 2 η + 16 j 3 η η j 3 ξ ξ + η 4 + 10 η 2 j 2 η + 2 η 2 32 j 3 η η j 2 ξ = 0
A more detailed treatment, including with torsional modes, was provided by Eringen and Șuhubi [22] (pp. 804–833), although it should be noted that there was an error in their solution for the (spherical) l = 0 breathing mode [23].
The eigenvalue equations shown above have multiple roots (Figure A1). The harmonic index indicates which root is being found, i.e., n = 1 indicates the first root, n = 2 indicates the second, etc.
Figure A1. Evaluation of eigenvalue equations for a sphere with R = 50 nm, Vl = 1920 m/s, and Vt = 960 m/s. Circles denote the roots. (a) Breathing mode (l = 0, Equation (A2)). (b) Dipolar mode (l = 1, Equation (A1) or (A4)). (c) Quadrupolar mode (l = 2, Equation (A1) or (A5)).
Figure A1. Evaluation of eigenvalue equations for a sphere with R = 50 nm, Vl = 1920 m/s, and Vt = 960 m/s. Circles denote the roots. (a) Breathing mode (l = 0, Equation (A2)). (b) Dipolar mode (l = 1, Equation (A1) or (A4)). (c) Quadrupolar mode (l = 2, Equation (A1) or (A5)).
Viruses 15 02110 g0a1
The eigenvalue equations should be solved numerically in order to determine the frequency of a specific vibrational mode. The python script available at github.com/gavinscode/sphere_resonance [19] (accessed on 5 August 2023) implements such a solution:
  • We defined a function E(f) as the eigenvalue Equations (A1) or (A2) (Figure A1).
  • E(f) was incrementally evaluated with increasing values of f (e.g., f1, f2, …) to determine when the function crossed zero, i.e., when E(f1) > 0 and E(f2) < 0, or vice versa.
  • When a zero-crossing occurred, we used a bounded minimiser from the SciPy package [24] to find the local minima of |E(f)| (where f1 < f < f2), which corresponded to the root of the eigenvalue equation. Note that there were local minima of |E(f)| which did not correspond to roots (Figure A1).
  • Solving for the frequency of harmonics greater than 1 was carried out sequentially, i.e., solving for n = 3 required initially finding the first (n = 1) and second (n = 2) roots.

References

  1. Manna, A.; De Forni, D.; Bartocci, M.; Pasculli, N.; Poddesu, B.; Lista, F.; De Santis, R.; Amatore, D.; Grilli, G.; Molinari, F.; et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443. [Google Scholar] [CrossRef] [PubMed]
  2. Lamb, H. On the Vibrations of an Elastic Sphere. Proc. Lond. Math. Soc. 1881, 13, 189–212. [Google Scholar] [CrossRef]
  3. Cantu, J.C.; Butterworth, J.W.; Mylacraine, K.S.; Ibey, B.L.; Gamboa, B.M.; Johnson, L.R.; Thomas, R.J.; Payne, J.A.; Roach, W.P.; Echchgadda, I. Evaluation of Inactivation of Bovine Coronavirus by Low-Level Radiofrequency Irradiation. Sci. Rep. 2023, 13, 9800. [Google Scholar] [CrossRef] [PubMed]
  4. Echchgadda, I.; Cantu, J.C.; Butterworth, J.; Gamboa, B.; Barnes, R.; Freeman, D.A.; Ruhr, F.A.; Williams, W.C.; Johnson, L.R.; Payne, J.; et al. Evaluation of Viral Inactivation on Dry Surface by High Peak Power Microwave (HPPM) Exposure. Bioelectromagnetics 2023, 44, 5–16. [Google Scholar] [CrossRef] [PubMed]
  5. Karaböce, B.; Hamid, R.; Saban, E.; Sözeri, H.; Danacı, E.; Böyük, A.A.; Durmuş, H.; Dorosinskiy, L.; Kızılbey, O.; Bilecik, S.; et al. Comparison of Disinfection of Phi6 Virus Bacteriophage on Fomites by Various Methods and a Prototype for IR Application. SSRN 2023. [Google Scholar] [CrossRef]
  6. Banting, H.; Goode, I.; Flores, C.E.G.; Colpitts, C.C.; Saavedra, C.E. Electromagnetic Deactivation Spectroscopy of Human Coronavirus 229E. Sci. Rep. 2023, 13, 8886. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, P.-J.; Huang, T.-W.; Chen, Y.-J.; Sun, C.-K. Implementation of a Coplanar-Waveguide Chip for the Measurement of EM Wave Absorption Spectrum of SARS-Cov-2 Virus. In Advances in Terahertz Biomedical Imaging and Spectroscopy; Zaytsev, K.I., Ponomarev, D.S., Skorobogatiy, M., Eds.; SPIE: Bellingham, WA, USA, 2022. [Google Scholar] [CrossRef]
  8. Hoff, B.W.; Cohick, Z.W.; Tilley, B.S.; Luginsland, J.W.; Revelli, D.; Cox, J.; Irshad, H.; Snider, A.; Arndt, A.; Ibey, B.L.; et al. Observed Reductions in the Infectivity of Bioaerosols Containing Bovine Coronavirus Under Repetitively Pulsed RF Exposure. IEEE Trans. Biomed. Eng. 2023, 70, 640–649. [Google Scholar] [CrossRef] [PubMed]
  9. Yang, S.-C.; Lin, H.-C.; Liu, T.-M.; Lu, J.-T.; Hung, W.-T.; Huang, Y.-R.; Tsai, Y.-C.; Kao, C.-L.; Chen, S.-Y.; Sun, C.-K. Efficient Structure Resonance Energy Transfer from Microwaves to Confined Acoustic Vibrations in Viruses. Sci. Rep. 2015, 5, 18030. [Google Scholar] [CrossRef] [PubMed]
  10. Lin, H.-C. Inactivation of Virus by Neutral Electrolyzed Water and Microwave Resonance. Master’s Thesis, National Taiwan University, Taipei, Taiwan, 2013. [Google Scholar] [CrossRef]
  11. Liu, T.-M.; Chen, H.-P.; Wang, L.-T.; Wang, J.-R.; Luo, T.-N.; Chen, Y.-J.; Liu, S.-I.; Sun, C.-K. Microwave Resonant Absorption of Viruses through Dipolar Coupling with Confined Acoustic Vibrations. Appl. Phys. Lett. 2009, 94, 043902. [Google Scholar] [CrossRef]
  12. Wang, P.-J.; Pang, Y.-H.; Huang, S.-Y.; Fang, J.-T.; Chang, S.-Y.; Shih, S.-R.; Huang, T.-W.; Chen, Y.-J.; Sun, C.-K. Microwave Resonant Absorption of SARS-CoV-2 Viruses. Sci. Rep. 2022, 12, 12596. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, T.-M.; Chen, H.-P.; Yeh, S.-C.; Wu, C.-Y.; Wang, C.-H.; Luo, T.-N.; Chen, Y.-J.; Liu, S.-I.; Sun, C.-K. Effects of Hydration Levels on the Bandwidth of Microwave Resonant Absorption Induced by Confined Acoustic Vibrations. Appl. Phys. Lett. 2009, 95, 173702. [Google Scholar] [CrossRef]
  14. Sun, C.-K.; Tsai, Y.-C.; Chen, Y.-J.E.; Liu, T.-M.; Chen, H.-Y.; Wang, H.-C.; Lo, C.-F. Resonant Dipolar Coupling of Microwaves with Confined Acoustic Vibrations in a Rod-Shaped Virus. Sci. Rep. 2017, 7, 4611. [Google Scholar] [CrossRef] [PubMed]
  15. Barbora, A.; Minnes, R. Targeted Antiviral Treatment Using Non-Ionizing Radiation Therapy for SARS-CoV-2 and Viral Pandemics Preparedness: Technique, Methods and Practical Notes for Clinical Application. PLoS ONE 2021, 16, e0251780. [Google Scholar] [CrossRef] [PubMed]
  16. Stephanidis, B.; Adichtchev, S.; Gouet, P.; McPherson, A.; Mermet, A. Elastic Properties of Viruses. Biophys. J. 2007, 93, 1354–1359. [Google Scholar] [CrossRef] [PubMed]
  17. Murray, D.B.; Saviot, L. Damping by Bulk and Shear Viscosity for Confined Acoustic Phonons of a Spherical Virus in Water. J. Phys. Conf. Ser. 2007, 92, 012036. [Google Scholar] [CrossRef]
  18. Arfken, G.B.; Weber, H.J.; Harris, F.E. Mathematical Methods for Physicists, 7th ed.; Academic Press: Cambridge, MA, USA, 2012; ISBN 978-0-12-384654-9. [Google Scholar]
  19. Taylor, G. Gavinscode/Sphere_resonance: Release for Zenodo. Available online: https://zenodo.org/record/8209172 (accessed on 5 August 2023).
  20. Murray, D.B.; Saviot, L. Phonons in an Inhomogeneous Continuum: Vibrations of an Embedded Nanoparticle. Phys. Rev. B 2004, 69, 094305. [Google Scholar] [CrossRef]
  21. Scheller, C.; Krebs, F.; Minkner, R.; Astner, I.; Gil-Moles, M.; Wätzig, H. Physicochemical Properties of SARS-CoV-2 for Drug Targeting, Virus Inactivation and Attenuation, Vaccine Formulation and Quality Control. Electrophoresis 2020, 41, 1137–1151. [Google Scholar] [CrossRef] [PubMed]
  22. Eringen, A.C.; Șuhubi, E.S. Elastodynamics; Academic Press: Cambridge, MA, USA, 1974; ISBN 978-0-12-240602-7. [Google Scholar]
  23. Saviot, L.; Murray, D.B.; Mermet, A.; Duval, E. Comment on “Estimate of the Vibrational Frequencies of Spherical Virus Particles”. Phys. Rev. E 2004, 69, 023901. [Google Scholar] [CrossRef] [PubMed]
  24. Virtanen, P.; Gommers, R.; Oliphant, T.E.; Haberland, M.; Reddy, T.; Cournapeau, D.; Burovski, E.; Peterson, P.; Weckesser, W.; Bright, J.; et al. SciPy 1.0: Fundamental Algorithms for Scientific Computing in Python. Nat. Methods 2020, 17, 261–272. [Google Scholar] [CrossRef] [PubMed]
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

Taylor, G.J.; Margueritat, J.; Saviot, L. Comment on Manna et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443. Viruses 2023, 15, 2110. https://doi.org/10.3390/v15102110

AMA Style

Taylor GJ, Margueritat J, Saviot L. Comment on Manna et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443. Viruses. 2023; 15(10):2110. https://doi.org/10.3390/v15102110

Chicago/Turabian Style

Taylor, Gavin J., Jérémie Margueritat, and Lucien Saviot. 2023. "Comment on Manna et al. SARS-CoV-2 Inactivation in Aerosol by Means of Radiated Microwaves. Viruses 2023, 15, 1443" Viruses 15, no. 10: 2110. https://doi.org/10.3390/v15102110

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