Next Article in Journal
A Review of the Image Classification Models Used for the Prediction of Bed Defects in the Selective Laser Sintering Process
Previous Article in Journal
Development of an Automated Downtime System for Small and Medium sized Enterprises
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Effect of Violet Laser Irradiation on the Optical Properties of Polyvinyl Alcohol/Methyl Orange Composite Thick Films: A Model for Medical Applications †

by
Sarah Maysam Tareq
1,
Nihal A. AbdulWahhab
1 and
Addnan H. Al-araji
2,*
1
Department of Physics, College of Science, University of Babylon, Babylon 5001, Iraq
2
Department of Medical Physics, Hilla University College, Babylon 5001, Iraq
*
Author to whom correspondence should be addressed.
Presented at the International Conference on Recent Advances in Science and Engineering, Dubai, United Arab Emirates, 4–5 October 2023.
Eng. Proc. 2023, 59(1), 236; https://doi.org/10.3390/engproc2023059236
Published: 27 February 2024
(This article belongs to the Proceedings of Eng. Proc., 2023, RAiSE-2023)

Abstract

:
This study investigates the impact of violet laser irradiation on the optical properties of thick films composed of polyvinyl alcohol (PVA), methyl orange (MO), and their composite (PVA/MO). Aimed at exploring potential medical applications, the films were synthesized through a casting process involving the dissolution of PVA and MO in distilled water. The optical properties, including absorbance spectra, energy gaps, and various optical constants, were meticulously measured before and after exposure to laser irradiation. The results revealed a notable decrease in the absorbance spectra and optical constants, along with an increase in the energy gaps, suggesting a structural modification induced by the laser treatment. These findings hold significance for the advancement of materials with customized optical features, potentially serving as a model for future developments in optoelectronic and photovoltaic devices. The research outcomes provide a foundation for the exploration of polymers and dyes in medical applications, particularly in the realms of non-invasive surgical procedures and simulations.

1. Introduction

Polymers, derived from the Greek words “poly”, meaning “many”, and “mere”, indicating “parts”, refer to high-molecular-mass molecules that consist of numerous repeating structural units [1]. Polyvinyl alcohol (PVA), a synthetic polymer, gained widespread use in the early twentieth century and has become a pivotal water-soluble plastic. Its inherent versatility allows it to easily blend with various substances, rendering it ideal for diverse applications. PVA has been employed across multiple sectors—industrial, commercial, medical, and food—for the production of lacquers, resins, surgical threads, and food packaging materials that are in direct contact with food [2]. The potential of PVA to undergo crosslinking through chemical or physical means, amalgamate with different polymers or copolymers, and graft or copolymerize with various monomers facilitates the creation of intelligent hydrogels. These hydrogels are notable for their stimuli-responsive properties, which can be fine-tuned. Moreover, PVA hydrogels exhibit excellent compatibility with skin, tissues, mucosa, and blood [3]. The utility of PVA extends to the simulation of soft tissue deformation, which can significantly enhance the effectiveness and precision of therapeutic interventions and minimally invasive surgical procedures. For these simulations, tissue-mimicking phantoms that emulate the mechanical properties of human or animal tissues are indispensable [4]. The structural formula of PVA is depicted in Figure 1 [5]. Dyes are characterized as organic compounds designed to color substrates through absorption, adsorption, reaction, or deposition with a certain level of permanence [6]. Methyl orange, an organic dye with the chemical formula ( CH 2 CH [ CH 2 CH ( CH 3 ) 2 ] ) n , is one such compound [7,8]. This dye is represented in Figure 2 below [9].
A composite material is defined as an assembly composed of two or more distinct substances, where the resulting properties are superior to those of the individual constituents [10,11,12,13]. These materials have been integral in addressing engineering challenges and are now widely utilized in various applications, including the automotive industry [14,15,16,17]. When laser radiation encounters a surface, it undergoes reflection, absorption, and transmission. The Beer–Lambert law quantitatively describes the absorption process as light traversing a medium [18]:
I = I 0 e α t
where I 0 and I denote the intensities of the incident and transmitted light, respectively, and α represents the absorption coefficient. The interaction between light and matter reveals the material’s optical properties. When the photon energy h ν is equal to or surpasses the band gap energy E g , it can excite a valence electron to the conduction band, generating an electron–hole pair. The threshold wavelength λ c for this interaction is given by [19]:
λ ( nm ) = h c E g = 1240 E g ( eV )
Direct photon transitions within the material are governed by [20]:
α h ν = B ( h ν E g ) 1 2
The optical characteristics of materials are those discovered when electromagnetic light strikes them [21]. The equation for reflectance, encompassing the sum of reflectance R, transmittance T, and absorbance A, is stated as [22]:
R + T + A = 1
The absorbance A is determined through [22]:
A = log I 0 I
This leads to the absorption coefficient α , calculated from the absorbance spectrum using [23]:
α = 2.303 A t
The extinction coefficient k is related to the absorption coefficient and the light’s wavelength by [24]:
k = α λ 4 π
The refractive index n, representing the ratio of light’s velocity in a vacuum to that within the material, can be calculated as [25]:
n = 4 R ( 1 R ) 2 k 2 + ( 1 + R ) ( 1 R )
The real and imaginary parts of the dielectric constants of thick films are determined by [26]:
ϵ r = n 2 k 2
ϵ i = 2 n k
The optical conductivity σ o p is influenced by the refractive index, speed of light, and absorption coefficient, as follows [19]:
σ o p = α n c 4 π
XRD is an invaluable non-destructive technique for characterizing materials, capable of discerning crystal structures, orientations, and sizes. The crystallographic planes are determined by Bragg’s law [27]:
2 d sin θ = n λ

2. Materials and Methods

Polyvinyl alcohol (PVA) and methyl orange (MO) powders, procured from Sigma-Aldrich Company (St. Louis, MI, USA), served as the primary materials in this study. A mass of 8 mg of PVA, with a molecular weight of 10,000 g/mol, and 7.5 mg of MO, with a molecular weight of 327.33 g/mol, were accurately weighed and dissolved in 7 mL of distilled water. The dissolution process was facilitated by heating the solution to 90 °C and maintaining continuous agitation with a magnetic stirrer for 10 min at room temperature, ensuring the complete solubility of both PVA and MO. Thereafter, the homogeneous solutions were cast into Petri dishes using the casting method and left to dry under ambient conditions, resulting in films with a nominal thickness of 3000 nm for each component, determined through a standardized gravimetric thickness measurement technique. The same protocol was employed to fabricate composite films of (PVA/MO), achieving a combined thickness of approximately 6000 nm under identical conditions. The films underwent irradiation by a 405 nm violet laser with an 80 mW power output for varying exposure times—10, 20, 30, and 40 min. The laser irradiation setup was calibrated to maintain a fixed distance between the laser source and the film surface, coupled with a heat sink to mitigate any thermal effects.
Post-irradiation, the optical properties of the films were quantified using a UV–Vis spectrophotometer (Model UV-1800 OA, Shimadzu, Kyoto, Japan), with a spectral range of 190–1200 nm. To ascertain the transmittance and absorbance, the films were analyzed using the same instrument settings both before and after laser exposure. Fourier transform infrared (FTIR) spectroscopy was also conducted to characterize the PVA powder. The FTIR spectrum was acquired using an IR Affinity-1 spectrometer (Shimadzu Company, Japan) with a detection range of 400–4000 cm−1. Data analysis, including the calculation of the optical properties, was executed with specially tailored Excel-based software, ensuring rigorous adherence to established analytical methods in spectrophotometry.

3. Results and Discussion

The Fourier transform infrared (FTIR) spectroscopy analysis of the PVA powder is depicted in Figure 3. Notable peaks were observed in the spectrum, corresponding to various functional groups. The peak at 1263.42 cm−1 can be attributed to the carbonyl (-C-O) wagged vibration, whereas the peaks at 1375.29 cm−1 and 1570.11 cm−1 can be ascribed to the (C-H) and (CH2) wagged vibrations, respectively. Furthermore, the presence of the carbonyl (−C=O−) group was confirmed by a peak at 1722.49 cm−1, which is indicative of the occurrence of a PVA reaction [28].
Figure 4 presents the FTIR spectrum for the MO powder. The strong peak detected at approximately 1280 cm−1 can be attributed to the C-S bond of the sulfonate group. Additionally, a weak band near 3300 cm−1 suggests the presence of an amine group. The significant peaks at around 1350 cm−1 are indicative of the C-N stretch of aromatic amine, which confirms the presence of the aromatic amine within the compound. The peak at around 1290 cm−1 denotes the S-O bond [29].
The X-ray diffraction (XRD) patterns for pure PVA are exhibited in Figure 5. The XRD pattern reveals a distinct diffraction band at 2 θ = 19 . 4 , characteristic of the semicrystalline nature of PVA. This semicrystallinity is likely due to the extensive intra-molecular hydrogen bonding within the PVA monomer units, as well as the inter-molecular hydrogen bonds among different monomer units [30]. The UV–Visible absorption spectra for the PVA, MO, and PVA/MO thick films, both before and after laser irradiation at varying times (0, 10, 20, 30, and 40 min), were examined.
PVA exhibited an absorption peak at 350 nm, MO at 471 nm, and the PVA/MO composite at around 480 nm, with no significant absorption peaks at longer wavelengths. With increasing irradiation times, a decrease in absorption was observed across all film thicknesses. This trend is depicted in Figure 6a, Figure 7a, and Figure 8a, while the absorption spectra as a function of irradiation time are shown in Figure 6b, Figure 7b, and Figure 8b.
Figure 9, Figure 10 and Figure 11 illustrate the energy gap calculations using Tauc’s method for PVA, MO, and PVA/MO, respectively. The squared product of the absorption coefficient and photon energy ( ( α h v ) 2 ) was plotted against the photon energy for various irradiation times. An increase in the violet laser irradiation time led to a corresponding increase in the direct energy gap, as evidenced in Figure 9b, Figure 10b, and Figure 11b. This phenomenon, marked by a decrease in the absorption coefficient, is a pivotal feature for photovoltaic applications.
The optical constants for the PVA, MO, and PVA/MO composites, including the absorption coefficient, refractive index, extinction coefficient, and dielectric constants, are summarized in Table 1, Table 2 and Table 3.

4. Conclusions

This investigation delineates the synthesis and characterization of thick films composed of polyvinyl alcohol (PVA), methyl orange (MO), and a PVA/MO composite, which were meticulously prepared by solubilizing measured quantities of PVA and MO powders in distilled water at a controlled temperature of 90 °C. The subsequent casting process resulted in uniform films with reproducible thicknesses. This research focused on alterations in the optical properties of these films due to violet laser irradiation. A systematic decrease in the absorbance spectra with increasing irradiation times was observed across all samples, indicating photo-induced modifications in the material structure. Concurrently, the photonic energy gaps were found to widen with increased laser exposure, suggesting alterations in the electronic band structure conducive to photovoltaic applications. Furthermore, a comprehensive analysis revealed that critical optical parameters, including the absorption coefficient, extinction coefficient, refractive index, and dielectric constants, uniformly decreased as a function of the irradiation time. This trend is indicative of the laser-induced densification of the material’s structure or a photochemical modification leading to a more ordered state, which has profound implications for enhancing the efficiency of optoelectronic devices. The outcomes of this study not only contribute to the fundamental understanding of polymer and dye interactions under laser irradiation but also pave the way for the development of novel materials with tailored optical properties. These materials have potential applications in a broad spectrum of fields, including photovoltaic systems, optical filters, and sensors, thereby holding promise for future technological advancements in material science.

Author Contributions

Methodology and writing—original draft preparation, A.H.A.-a.; validation, S.M.T.; formal analysis, N.A.A.; investigation, S.M.T.; resources, N.A.A.; data curation, S.M.T.; writing—review and editing, A.H.A.-a. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data used in the experiment have been made available in the present article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ebewele, R.O. Polymer Science and Technology; CRC Press: New York, NY, USA, 1995; Volume 16, p. 3. [Google Scholar]
  2. Nagarkar, R.; Patel, J. Polyvinyl Alcohol: A Comprehensive Study. Acta Sci. Pharm. Sci. 2019, 3, 34–44. [Google Scholar]
  3. Patachia, A.P.; Valente, A.J.M. (Eds.) Poly (vinyl alcohol) [PVA]-Based Polymer Membranes; Nova Science Publishers: New York, NY, USA, 2009. [Google Scholar]
  4. Jiang, S.; Liu, S.; Feng, W. PVA hydrogel properties for biomedical application. J. Mech. Behav. Biomed. Mater. 2011, 4, 1228–1233. [Google Scholar] [CrossRef]
  5. Demerlis, C.C.; Schoneker, D.R. Review of the oral toxicity of polyvinyl alcohol (PVA). Food Chem. Toxicol. 2003, 41, 319–326. [Google Scholar] [CrossRef]
  6. Giri, M.; Singh, D.; Lal, J.; Jaggi, N.; Singh, N.; Jaiswal, R.M.P. Absorption and Fluorescence Spectra of Methyl Orange in Aqueous Solutions. Atti Fond. Giorgio Ronchi ANNO LXVII 2012, 2, 255. [Google Scholar]
  7. Al-kadhemy, M.F.H.; Saeed, A.A.; Kadhum, F.J.; Mazloum, S.A.; Aied, H.K. The effect of (He–Ne) laser irradiation on the optical properties of methyl orange doped PVA films. J. Radiat. Res. Appl. Sci. 2014, 7, 371–375. [Google Scholar] [CrossRef]
  8. Bujdák, J. Controversial Issues Related to Dye Adsorption on Clay Minerals: A Critical Review. Molecules 2023, 28, 6951. [Google Scholar] [CrossRef]
  9. Mohammadi, N.; Khani, H.; Gupta, V.K.; Amereh, E.; Agarwal, S. Adsorption process of methyl orange dye onto mesoporous carbon material-kinetic and thermodynamic studies. J. Colloid Interface Sci. 2011, 362, 457–462. [Google Scholar] [CrossRef]
  10. Mohammed, H.S. Study the Effect of Immobilization on Electrostatic Potential for Methyl Orange Dye. Muthanna J. Pure Sci. 2018, 5, 38–43. [Google Scholar] [CrossRef]
  11. Bhat, R.; Mohan, N.; Sharma, S.; Rao, S. Influence of Seawater Absorption on the Hardness of Glass Fiber/Polyester Composite. J. Comput. Mech. Manag. 2022, 1, 1–10. [Google Scholar] [CrossRef]
  12. Singh, H.; Singh, K.; Vardhan, S. Enhancing Aluminum Matrix Composites with Hexagonal Boron Nitride (hBN) Particulates: A Comprehensive Review. J. Comput. Mech. Manag. 2023, 2, 5. [Google Scholar] [CrossRef]
  13. Srikanth, V.; Kowshik, S.; Narasimha, D.; Patil, S.; Samanth, K.; Rathee, U. Finite Element Modelling and Analysis of Fiber Reinforced Concrete under Tensile and Flexural Loading. J. Comput. Mech. Manag. 2022, 1, 11–17. [Google Scholar] [CrossRef]
  14. Chawla, K.K. Composite Materials; Springer: New York, NY, USA, 2012. [Google Scholar] [CrossRef]
  15. Naik, N.; Suresh, P.; Yadav, S.; Nisha, M.P.; Arias-Gonzáles, J.L.; Cotrina-Aliaga, J.C.; Bhat, R.; Jalageri, M.D.; Kaushik, Y.; Kunjibettu, A.B. A Review on Composite Materials for Energy Harvesting in Electric Vehicles. Energies 2023, 16, 3348. [Google Scholar] [CrossRef]
  16. Hiremath, P.; Viswamurthy, S.R.; Shettar, M.; Naik, N.; Kowshik, S. Damage Tolerance of a Stiffened Composite Panel with an Access Cutout under Fatigue Loading and Validation Using FEM Analysis and Digital Image Correlation. Fibers 2022, 10, 105. [Google Scholar] [CrossRef]
  17. Mohan, N.; Sharma, S.; Bhat, R. A Comprehensive Study of Glass Fibre Reinforced Polymer (GFRP) Drilling. Int. J. Mech. Prod. Eng. Res. Dev. 2019, 9, 1–10. [Google Scholar]
  18. Nowakowski, K.A. Laser Beam Interaction with Materials for Microscale Applications. Doctoral Dissertation, Worcester Polytechnic Institute, Worcester, MA, USA, 2005. [Google Scholar]
  19. Hammood, A.; Al-Aarajiy, M. Preparation and Characterization of NiPc/Si Organic Solar Cell; University of Baghdad: Baghdad, Iraq, 2014. [Google Scholar]
  20. Burns, G. Solid State Physics; Cambridge University Press: New York, NY, USA, 1985. [Google Scholar]
  21. Hasan, M.; Khaleel, N. Study the Linear and Nonlinear Optical Properties for Methylene Blue Dye Doped SiO2 Nanoparticles; University of Babylon: Babylon, Iraq, 2022. [Google Scholar]
  22. Ejam, A.A.; Wahhab, N.A.A. Concentration effect on the optical properties of laser irradiation copper phthalocyanine (CuPc blue) solution. NeuroQuantology 2022, 20, 1–15. [Google Scholar] [CrossRef]
  23. AbdulWahhab, N.A. Optical properties of SnO2 thin films prepared by pulsed laser deposition technique. J. Opt. 2020, 49, 41–47. [Google Scholar] [CrossRef]
  24. Nasir, E.M.; Hussein, M.T.; Al-Aarajiy, A.H. Investigation of Nickel Phthalocyanine Thin Films for Solar Cell Applications. Adv. Mater. Phys. Chem. 2019, 9, 158–173. [Google Scholar] [CrossRef]
  25. Abdullah, O.G.; Saber, D.R. Optical absorption of Polyvinyl alcohol films doped with Nickel Chloride. Appl. Mech. Mater. 2012, 110–116, 177–182. [Google Scholar] [CrossRef]
  26. Habubi, N.F.; Chiad, S.S. Optical Properties of Doped Polymers. Diyala J. Pure Sci. 2017, 7, 153–161. [Google Scholar]
  27. Sólyom, J. Fundamentals of the Physics of Solids, Volume 1 Structure and Dynamics; Springer: New York, NY, USA, 2007. [Google Scholar]
  28. Azim-Araghi, M.E.; Krier, A. Optical characterization of chloroaluminium phthalocyanine (ClAlPc) thin films. Pure Appl. Opt. J. Eur. Opt. Soc. 1997, 6, 443. [Google Scholar] [CrossRef]
  29. Thampraphaphon, B.; Phosri, C.; Pisutpaisal, N.; Thamvithayakorn, P.; Chotelersak, K.; Sarp, S.; Suwannasai, N. High Potential Decolourisation of Textile Dyes from Wastewater by Manganese Peroxidase Production of Newly Immobilised Trametes hirsuta PW17-41 and FTIR Analysis. Microorganisms. 2022, 10, 992. [Google Scholar] [CrossRef] [PubMed]
  30. Mahendia, S.; Tomar, A.K.; Chahal, R.P.; Goyal, P.; Kumar, S. Optical and structural properties of poly(vinyl alcohol) films embedded with citrate-stabilized gold nanoparticles. J. Phys. D Appl. Phys. 2011, 44, 20. [Google Scholar] [CrossRef]
Figure 1. Structural formula of PVA: (A) Partly hydrolyzed; (B) Completely hydrolyzed.
Figure 1. Structural formula of PVA: (A) Partly hydrolyzed; (B) Completely hydrolyzed.
Engproc 59 00236 g001
Figure 2. Structure of methyl orange dye.
Figure 2. Structure of methyl orange dye.
Engproc 59 00236 g002
Figure 3. FTIR spectrum for PVA powder.
Figure 3. FTIR spectrum for PVA powder.
Engproc 59 00236 g003
Figure 4. FTIR spectrum for MO powder.
Figure 4. FTIR spectrum for MO powder.
Engproc 59 00236 g004
Figure 5. XRD pattern for PVA powder.
Figure 5. XRD pattern for PVA powder.
Engproc 59 00236 g005
Figure 6. PVA absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of PVA for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for PVA at a 350 nm wavelength.
Figure 6. PVA absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of PVA for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for PVA at a 350 nm wavelength.
Engproc 59 00236 g006
Figure 7. MO absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of MO for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for MO at a 471 nm wavelength.
Figure 7. MO absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of MO for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for MO at a 471 nm wavelength.
Engproc 59 00236 g007
Figure 8. PVA/MO absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of PVA/MO for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for PVA/MO at a 480 nm wavelength.
Figure 8. PVA/MO absorbance spectra and absorbance as a function of irradiation time: (a) Absorbance spectra of PVA/MO for different laser exposure times (0, 10, 20, 30, and 40 min); (b) Absorbance as a function of irradiation time for PVA/MO at a 480 nm wavelength.
Engproc 59 00236 g008
Figure 9. PVA energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for PVA at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for PVA at a 350 nm wavelength.
Figure 9. PVA energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for PVA at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for PVA at a 350 nm wavelength.
Engproc 59 00236 g009
Figure 10. MO energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for MO at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for MO at a 471 nm wavelength.
Figure 10. MO energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for MO at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for MO at a 471 nm wavelength.
Engproc 59 00236 g010
Figure 11. PVA/MO energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for PVA/MO at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for PVA/MO at a 480 nm wavelength.
Figure 11. PVA/MO energy gap and direct energy gap as a function of irradiation time: (a) Energy gap ( α h v ) 2 for PVA/MO at different laser irradiation times (0, 10, 20, 30, and 40 min); (b) Direct energy gap as a function of irradiation time for PVA/MO at a 480 nm wavelength.
Engproc 59 00236 g011
Table 1. The optical constants of PVA thick films at 350 nm.
Table 1. The optical constants of PVA thick films at 350 nm.
Time (min) α × 10 4 kn ε real ε imag σ op × 10 14 E g (eV)
(cm−1) (s−1)
00.1070.00291.9413.7680.0110.0493.95
100.0930.00251.8403.3880.0090.0414.01
200.0820.00221.7553.0820.0080.0344.07
300.0720.00201.6722.7960.0060.0284.09
400.0610.00171.5802.4990.0050.0234.12
Table 2. The optical constants of MO thick films at 471 nm.
Table 2. The optical constants of MO thick films at 471 nm.
Time (min) α × 10 4 kn ε real ε imag σ op × 10 14 E g (eV)
(cm−1) (s−1)
00.0870.00321.7973.2310.01180.03772.7
100.0840.00311.7733.1450.01120.03592.9
200.0800.00301.7403.0290.01050.03353.0
300.0610.00231.5832.5070.00730.02333.1
400.0560.00211.5422.3780.00650.02093.2
Table 3. The optical constants of PVA/MO composite thick films at 480 nm.
Table 3. The optical constants of PVA/MO composite thick films at 480 nm.
Time (min) α × 10 4 kn ε real ε imag σ op × 10 14 E g (eV)
(cm−1) (s−1)
00.0540.00201.9533.8160.00810.02552.4
100.0380.00141.7182.9510.00510.01592.5
200.0310.00121.5972.5530.00380.01212.6
300.0260.00101.5152.2950.00310.00972.7
400.0230.00081.4462.0920.00250.00792.8
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

Tareq, S.M.; AbdulWahhab, N.A.; Al-araji, A.H. Effect of Violet Laser Irradiation on the Optical Properties of Polyvinyl Alcohol/Methyl Orange Composite Thick Films: A Model for Medical Applications. Eng. Proc. 2023, 59, 236. https://doi.org/10.3390/engproc2023059236

AMA Style

Tareq SM, AbdulWahhab NA, Al-araji AH. Effect of Violet Laser Irradiation on the Optical Properties of Polyvinyl Alcohol/Methyl Orange Composite Thick Films: A Model for Medical Applications. Engineering Proceedings. 2023; 59(1):236. https://doi.org/10.3390/engproc2023059236

Chicago/Turabian Style

Tareq, Sarah Maysam, Nihal A. AbdulWahhab, and Addnan H. Al-araji. 2023. "Effect of Violet Laser Irradiation on the Optical Properties of Polyvinyl Alcohol/Methyl Orange Composite Thick Films: A Model for Medical Applications" Engineering Proceedings 59, no. 1: 236. https://doi.org/10.3390/engproc2023059236

Article Metrics

Back to TopTop