Next Article in Journal
A Ferrofluid with High Specific Absorption Rate Prepared in a Single Step Using a Biopolymer
Next Article in Special Issue
Sunlike White Light-Emitting Diodes Based on Rare-Earth-Free Luminescent Materials
Previous Article in Journal
Effects of Moisture on NHCapture Using Activated Carbon and Acidic Porous Polymer Modified by Impregnation with H3PO4: Sorbent Material Characterized by Synchrotron XRPD and FT-IR
Previous Article in Special Issue
Advanced Photonic Thin Films for Solar Irradiation Tuneability Oriented to Greenhouse Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Optical Characterizations of Cadmium Chalcogenide Layers on Polyamide Formed Using Monotelluropentathionic Acid

Faculty of Chemical Technology, Department of Physical and Inorganic Chemistry, Kaunas University of Technology, 50254 Kaunas, Lithuania
*
Author to whom correspondence should be addressed.
Materials 2022, 15(3), 787; https://doi.org/10.3390/ma15030787
Submission received: 14 December 2021 / Revised: 12 January 2022 / Accepted: 18 January 2022 / Published: 20 January 2022
(This article belongs to the Special Issue Advanced Materials for Optical Applications and Devices)

Abstract

:
Mixed cadmium tellurides–cadmium sulfide thin layers were formed on the polyamide PA 6. Monotelluropentathionic acid (H2TeS4O6) was used as a precursor of tellurium and sulfur. A low-temperature, nontoxic, and cost-effective SILAR method was applied. Cadmium telluride (CdTe) and sulfide (CdS) layers were formed through the consecutive reactions of sorbed/diffused chalcogens species from telluropentathionate anion (TeS4O62−) with functional groups of polyamide and alkaline cadmium sulfate. The pseudo-second-order rate and Elovich kinetic models were the best fit to quantify an uptake of chalcogens and cadmium on PA 6. The effects of chalcogens and Cd on the structure and optical properties of PA 6 were characterized using UV-Vis and IR spectra. The clear changes of these properties depended on the concentration and exposure time in the precursor solutions. Fourier transform infrared spectroscopy and ultraviolet-visible spectroscopy were applied in order to evaluate the effect of the chalcogen species on the changes in structure of polyamide 6 films, depending on the exposure time in the solution of the chalcogens precursor and its concentration. The optical bandgap energy of the formed layers was found to be in the order of 1.52–2.36 eV. Studies by scanning electron microscopy and atomic force microscopy reveal that the diameter of the average grain is approximately 30 nm. The grains are conical in shape and unevenly distributed all over the surface of the substrate.

Graphical Abstract

1. Introduction

Compounds of the II–VI group of elements or semiconductors based on these compounds are important luminescent, lasing, and optoelectronic materials [1]. Due to the successful synthesis of semiconductor compounds of type II–VI, these materials can be widely used for mechanical and optoelectronic purposes; therefore, further study of such semiconductors is extremely important [1,2,3]. Due to its high optical absorption coefficient (>104 cm−1) and optimal bandgap (1.5 eV), CdTe is an efficient photoconverter [2]. Therefore, it is one of the main widely used photovoltaic materials for the successful development of high-efficiency solar cells and modules.
Cadmium chalcogenides CdE (E = S, Te), e.g., cadmium sulfide, and telluride thin films are used in laser windows and photo-electric cells, photothermal conversion, and solar cells, etc. [3,4]. With the aim of low-cost solar cells having thin films, common preliminary materials are employed, which are used in the semiconductor industry [5,6]. CdTe and CdS are typical layered semiconductors with almost complete chemical bonds inside the layer and, because of this, they have a low density of dangling bonds on the surface [7,8] and mechanical weakness due to the weak Van der Waals forces between layers. It has also been found that the solar cell efficiency correlated with the chemical and morphological properties of the surfaces [9].
A wide range of methods have been applied to produce cadmium chalcogenide thin layers, including electrodeposition [10], high-frequency magnetron sputtering [11], spray pyrolysis [12], thermal co-evaporation [13], the thermal decomposition of cadmium organic precursors [14], vapor deposition [15], electron beam evaporation [16], and others. Unfortunately, most of these techniques require special equipment and conditions such as high temperature, vacuum, and conductive substrates, and they usually consist of many stages, making them unsuitable and costly for the manufacture of the majority of polymers.
Only chemical methods used in the formation process of cadmium chalcogenide thin layers allow the development of good quality thin layers through the application of low temperature conditions. Layers of chalcogenides of metals of the I–II B groups on the substrate of an amphiphilic polyamide 6 (PA 6) were successfully obtained by the wet chemical method of adsorption–diffusion technology [17,18,19,20,21]. Very similar to this method, the successive ionic layer adsorption and reaction (SILAR) technique of the wet chemical methods is widely used to form these layers on a glass substrate [22,23,24]. In any case, due to its efficiency and economy, the SILAR method is one of the most promising for obtaining high-quality thin layers of chalcogenides on polymers for photovoltaic applications. In this method, the formation of metal chalcogenide layers takes place when the polymer substrate is exposed to a solution of the chalcogens precursor and further in a solution of metal ions with a suitable complexing agent. The production of thin-layer solar cells on polymer substrates or flexible metal foils offers several advantages for both space and terrestrial applications. The thickness of the solar cell is only 5 µm, while the thickness of the flexible substrate is 5–10 µm, so the weight savings compared to glass substrates are significant.
This work evaluated more closely what effect chalcogens have on the properties of semiconductors, along with their suitability for photovoltaic application. The selection of suitable precursors and their studies illustrate how chalcogens’ compounds could be used in electroactive materials. The preparation of chalcogen layers using chalcopolythionic acids, such as monoselenopentathionic H2SeS4O6 and monotelluropentathionic H2TeS4O6, has caught the attention of researchers [20,25,26]. These precursors are more beneficial than conventional sources due to their higher stability. It was determined that, under our experimental conditions, the solution of monotelluropentathionic acid was more stable than the alkaline metal salt solution of a similar concentration [26] used in our previous experiment.

2. Materials and Methods

2.1. Materials

Tellurous acid (98% purity), sulfuric acid (95%), acetic acid (99%), and ammonia water (25%) were acquired from Labochema (Vilnius, Lithuania) Na2S2O3 × 5H2O (99.5% purity); barium perchlorate (97% purity) and CdSO4 × 8/3H2O (99% purity) were purchased from SigmaAldrich, Saint Louis, MO, USA. Film of polyamide 6 (PA 6) (specification TU 6-05-1775-76) and grade PK-4 (Nylon 6) was used as a substrate. The main characteristics of Polyamide 6 (Nylon 6) are presented in Table 1.

2.2. Cadmium Chalcogenides Layer Formation Materials

Films of polyamide 6, 15 mm × 70 mm in size and 70 μm in thickness, were used. The porosity of the polymer was determined by the β-method on a Quantasorb (Osaka, Japan). The pore size of PA 6 was found to be much less than 1.5 nm, indicating that it is practically non-porous [27]. Prior to chalcogenization, the samples of polyamide were boiled in distilled water at 100 °C for 2 h to remove the remaining un-polymerized monomer residues and dried in the desiccator over dehydrated CaCl2 for 24 h. Then, the samples of PA 6 were used in the analysis and in further experiments. The formation scheme of cadmium chalcogenide layers is shown in Figure 1.
Firstly, the samples of polyamide film were chalcogenized at 25 °C in aqueous solutions of 0.05 and 0.1 mol/L monotelluropentathionic acid. The experimental conditions are shown in Table 2.
Sodium telluropentathionate, Na2TeS4O6, was synthesized according to the procedure [28]:
H2TeO3 + 4Na2S2O3 + 4CH3COOH → Na2TeS4O6 + Na2S4O6 + 4CH3COONa + 3H2O
Then, BaTeS4O6 was produced using barium perchlorate:
Na2TeS4O6 + Ba(ClO4)2 → BaTeS4O6 + 2NaClO4
H2TeS4O6 acid was isolated from synthesized barium salt in the reaction of BaSO4 precipitation with the solution of 3 mol/L sulphuric acid:
BaTeS4O6 + H2SO4 → H2TeS4O6 + BaSO4
Next, to form mixed cadmium chalcogenide layers, chalcogenized samples of PA 6 were treated with an alkaline (pH was found to be 11.5) cadmium salt solution at 80 °C for 5 min. An amount of 0.1 mol/L CdSO4 solution was made from crystalline CdSO4 × 8/3H2O; the pH of this solution was adjusted using a solution of ammonia. A complex cation was formed according to the reaction (4):
Cd2+ + 4NH3 × H2O → [Cd(H2O)2(NH3)4]2+ + 2H2O
After treatment in a cadmium salt solution, the samples were washed in distilled water, dried in a desiccator over CaCl2 for 24 h, and analyzed.

2.3. Instrumental Analysis

The UV-Vis (190–800 nm) and IR (400–1400 cm−1) spectra were recorded by a spectrophotometer Spectronic® GenesysTM 8 (Spectronic instruments, Runcorn, UK) and spectrometer Perkin Elmer GX system FTIR (Thermo Fisher Scientific, Waltham, MA, USA), respectively.
A D8 Advance diffractometer (Bruker AXS, Karlsruhe, Germany) was used for X-ray diffraction analysis of the deposited layers of cadmium chalcogenide. Operating voltage and current in the tubes of the device were 40 kV and 40 mA. To record diffraction patterns in the Bragg–Brentano geometry, a fast counting one-dimensional Bruker LynxEye detector (Bruker AXS, Karlsruhe, Germany) based on silicon strip technology was used. The X-ray was filtered through a 0.02 mm Ni filter to suppress Cu-k alpha radiation, and the samples were scanned in the 2θ = 3–70° range at a scan rate of 6° 1/min. The diffractometer is supplied together with the software package DIFFRAC.SUITETM. X-ray diffractograms of deposited layers were processed using the software packages Crystallographica Search-Match v. 2.1, ConvX v.1.0 and Microsoft Office Excel 2013.
To obtain infrared spectra in the range of 400–1400 cm−1, a Perkin-Elmer FTIR Spectrum GX spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) was used.
The thickness of the layers was measured using the scanning electron microscopy (SEM) images of the cross-section. A sample cross section was prepared by cutting the layers of CdxSy–CdxTeyon PA 6 with a sharp razorblade and was mounted for imaging using carbon tape. SEM imaging was performed with a Scanning Electron Microscope Quanta 200 FEG (FEI, Eindhoven, The Netherlands). The microscope was also equipped with an Energy Dispersive X-ray Spectrometer (EDS) detector XFlash 4030 (Bruker AXS Microanalysis, Berlin, Germany).
The quantitative microscopic studies of the surface roughness and the morphology of the formed layers of cadmium chalcogenide on the PA 6 were carried out using a QUESANT QScope-250 atomic force microscope (AFM) (Quesant Instrument Corporation, Agoura Hills, CA, USA). The dry samples were investigated in contact mode using commercial (Nano Technology Instruments—Europe BV) Si cantilevers CSG10 series with a force constant k = 0.2 Nm−1 and tip curvature of 10 nm. The images were analyzed, and image metrology was made using SPIP (Image Metrology, A/S, Lyngby, Denmark) and Quesant Corp. software v. 2.3.
A Perkin-Elmer 503 atomic absorption spectrometer was used for determination of the total amount of cadmium (λ = 228.8 nm) and tellurium (λ = 214.3 nm) in the cadmium chalcogenides layers. The sensitivity of atomic adsorption spectroscopic determination for Te and Cd, respectively, is 1.0 μg/cm3 and 0.025 μg/cm3 for 1% absorption.
The concentration of sulfur in PA 6 as sulfate was measured using the turbidimetric method [29] with a photoelectric colorimeter KFK-4 (Zagorsk Optical-Mechanical Plant, Sergiyev Posad, Russia) at λ = 400.0 nm. SO42− in the concentration range of 1–15 mg/dm3 can be determined by reaction with BaCl2 in the form of BaSO4. The deviation in the concentration interval of 5−10 mg/dm3 is 8%.

2.4. Modeling the Kinetics of Sorption

In order to quantify the extent of uptake in adsorption kinetics, six simple adsorption reaction and adsorption diffusion models were tested. Lagergren’s pseudo first-order rate equation is expressed as follows [30]:
l n ( q e q t ) = l n q e k 1 t
where qt is the adsorption capacity (μmol/cm2) at time t (h), qe is the adsorption capacity (μmol/cm2) at equilibrium time (48 h), and k1 is the rate constant of the pseudo first–order adsorbtion (h−1). The values of k1 and qe can be obtained from the intercept and slope of the plot of ln(qe qt) versus t.
A pseudo second-order adsorbtion kinetic rate equation (Ho’s equation) is [31]:
t q t = 1 k 2 q e 2 + t q e
where k2 is the rate constant of the pseudo second-order adsorbtion (cm2/μmol·h). The values of k2 and qe were obtained from the intercept and slope of the plot of t·qt1 versus t.
The integrated form of Elovich’s equation becomes [32]:
q t = 1 b l n ( t + t 0 ) 1 b l n t 0
t 0 = 1 a b
where a (μmol/cm2·h) is the initial adsorption rate. The constant a is regarded as the initial rate because (dqt/dt) approaches a when qt approaches 0 [33]; b (cm2/μmol) is the desorption constant. With the assumption of t >> t0 (this will be justified later), this is simplified as:
q t = 1 b l n ( a b ) + 1 b l n ( t )
The values of a and b can be obtained from the slope and intercept of the plot of qt versus lnt. The film diffusion mass transfer rate equation presented by Boyd et al. [34] is:
l n ( 1 q t q e ) = R t
where R is the liquid film diffusion constant (h−1). A plot of ln(1 − qt/qe) versus t should be a straight line with a slope −R if the film diffusion is the rate limiting step.
The Weber–Morris intraparticle diffusion model is [35]:
q t = k i n t t
where kint is intraparticle diffusion rate constant (μmol/cm2·h−1/2). A plot of qt versus t1/2 should be a straight line with a slope kint when the intraparticle diffusion is a rate-limiting step. Dumwald–Wagner proposed another intraparticle diffusion model as [36]:
l n ( 1 ( q t q e ) 2 ) = K t
where K is the Dumwald–Wagner rate constant of adsorption (h−1). A plot of ln(1 − (qt/qe)2) versus t should be linear, and the rate constant K can be obtained from the slope.
The experimental qt values are first processed by linearized equations to determine the model parameters, and, if r2 > 0.95, the determined values were used to reconstruct the kinetic equations. The kinetic curves show an overlap of experimental results (points). The linear correlation coefficients (r2) showed the correspondence of the experimental data to the linearized forms of the kinetic equations. The average percentage errors (APE) indicate the agreement between the experimental and predicted adsorption capacity values used to construct the kinetic curves [37]:
A P E = 100 n i = 1 n | q e x p e r i m e n t a l q p r e d i c t e d q e x p e r i m e n t a l |
where n is the number of experimental data.

2.5. Optical Property Measurements

The classical relation for the near-edge optical absorption of semiconductors was used to analyze the absorption data [38]:
α h ν = B ( h ν E g ) m  
where α is absorption coefficient, is photon energy, Eg is bandgap, and B is constant. In the equation, m can have values of 1/2, 2, 3/2, and 3 for allowed direct, allowed indirect, forbidden direct, and forbidden indirect transitions, respectively.
A dependence between the bandgap and wavelength, equivalent to the bandgap energy, λg, can be derived [39]:
E g = 1239.83 λ g  
λg is determined by extrapolating the linear portion of the curve on the abscissa:
( A λ ) 1 / m = f ( 1 λ )  
Because CdTe and CdS are direct bandgap semiconductors, an m value of 0.5 was used.
The absorption coefficient and the photon energy are related by the Urbach energy:
α = α 0 e h ν / E U  
where α0 is constant and EU is Urbach energy. Urbach energy can be calculated using a tangent of the angle of the linear portion of the curve lnA = λ−1, according to EU∙tgθ = 1239.83.

3. Results and Discussions

3.1. Chalcogens’ and Cadmium Adsorbtions Kinetics

The kinetic curves (Figure 2) were determined by changing the initial concentration of the precursor in 0.05 and 0.1 mol/L of the H2TeS4O6 solutions and using an exposure time of 48 h, which was found sufficient to reach saturation. The amount of adsorbed chalcogens containing species depends on precursor concentration and duration of treatment. An increase in the precursor concentration from 0.05 to 0.1 mol/L leads to a S content of approximately 1.3–2.1 µmol/cm2 of PA 6 (Figure 2, curves S/S1 and S/S2) and, respectively, the same Te of approximately 0.672–0.934 µmol/cm2 of PA 6 (Figure 2, curves Te/S1 and Te/S2) after 48 h of chalcogenization. In addition, the content of S and Te increases with the increase of exposure time. The concentration of S and Te reaches the saturation concentration, respectively, of 2.1 µmol/cm2 and 0.934 µmol/cm2 after 48 h of chalcogenation at 25 °C.
The cadmium adsorption kinetics on polyamide was investigated when the chalcogens and Cd2+ were adsorbed successively. The adsorption curves of Cd on the surface of chalcogenized PA (Figure 2, curves Cd/S1* and Cd/S2*) show that the adsorption of Cd increased as the concentration of chalcogens with species increased. Accordingly, an exposure for 48 h in 0.05 and 0.1 mol/L precursor solution yields a Cd content in PA 6, respectively, of 0.140 µmol/cm2 and 0.176 µmol/cm2. However, an adsorption of Cd ions on PA is not preferred in comparison with that of chalcogens.
The pseudo second-order rate kinetic model and Elovich‘s equation have been applied to describe the feasibility of chalcogens having species and cadmium sorption onto PA 6, which fitted best the experimental results (minimum APE) (Table 3). The maximum adsorption capacity increased slightly as the concentration increases in all cases. The molar ratio of Cd:Te:S after 48 h of chalcogenization is 1:4.8:9.3 for 0.05 mol/dm3 and, respectively, 1:5.3:11.9 for 0.1 mol/dm3 initial precursor concentration. The adsorption rate constants, the initial adsorption rate, and the desorption constants were found to increase with decreased concentration of Te and Cd, except for S, where they decreased.
The kinetics’ parameters obtained (e.g., best-fitted kinetic model; the pseudo-second-order rate constants; and the amount of Te, S, and Cd) demonstrate how they may be used in the formation of cadmium chalcogenides’ layers on PA 6 of the desirable phase composition and optical properties.

3.2. UV-Vis and FTIR Spectroscopies

UV-Vis spectra of chalcogenized PA 6 samples and the spectra of those after treatment in Cd2+ salt solution (with compensation of an absorption PA 6) are shown in Figure 3.
In UV-Vis spectra of chalcogenized polyamide, the absorption maxima, were observed at λ = 285–300 nm. Nevertheless, their maximum peak values are shifted towards the shorter wavelengths of the spectrum in comparison to the maximum values of UV-Vis spectra, 310–315 nm of the solutions of H2TeS4O6. All the layers of chalcogens having particles on PA 6 films have very low absorbance near the ultra-violet region of the spectrum (Figure 3, curves S1-2 and S1-24). However, percentage absorption (>90%) of chalcogens having particles containing layers on PA 6 was maximum at λ ~300 nm. With an increase in the exposure time from 2 to 24 h in solutions of the same precursor concentration, the intensity of the maximum of the absorption spectra increased approximately two-fold (Figure 3, curves S1-2 and S1-24, S2-2 and S2-24), and the maximum peaks are shifted towards the longer wavelength region. The intensity of the spectrum of PA 6 treated in 0.05 mol/L. H2TeS4O6 varies from 0.928 to 1.859 absorption units (a.u.) and, respectively, from 1.277 to 2.635 a.u. of the same treated in 0.1 mol/L H2TeS4O6. The concentration of the precursor also increases the effect of increasing absorption at the maximum wavelength; the clear shift to the direction of longer wavelength, e.g., 290–300 nm, was observed in the spectra of PA samples treated in 0.1 mol/L H2TeS4O6 in comparison with the wavelength region 285–295 nm of the spectra of the PA samples treated in 0.05 mol/L H2TeS4O6. A variety of new peaks occurs in the range 315 to 600 nm in the layers with cadmium chalcogenides compared to peaks before interaction with Cd2+ (Figure 3, curves S1*-2 and S1*-24, S2*-2 and S2*-24). These new broad and unclear peaks occur in the spectra of PA chalcogenized in the 0.05 mol/L H2TeS4O6 solution at 315–470 nm and, respectively, at 320–560 nm after chalogenization in 0.1 mol/L H2TeS4O6. Only the peaks in the 315–325 nm regions are sharper (Figure 3, curves S1*-2 and S1*-24, S2*-2 and S2*-24). These peaks are higher than the corresponding peaks of PA 6 before the interaction with cadmium, indicating the existence of cadmium chalcogenides. To describe the formation of cadmium chalcogenide in general terms, the formula CdTe–CdS is used. It is assumed that these new peaks show cadmium telluride or mixed cadmium telluride-sulfide, which occurred after treating with Cd2+. Some experiments are needed to clarify the attribution of the new UV-Vis absorption bands. Some of the methods applied provide information, but the data interpretation on the PA surface is complicated. The next experiment will focus on a mechanism in an aqueous solution and is explained in reactions on polyamide.
The spectra of optical absorption of the chalcogens species having layers on PA 6 showed that they have high absorbance (≥104 cm−1) in the UV-Vis region, with a gradual decrease towards the NIR region of the electromagnetic spectrum indicating direct bandgap transition.
According to literature data [40] the bandgap values are between 1.517 eV (CdTe) and 2.41 eV (CdS); see Table 4. The bandgap values of cadmium chalcogenide layers with different stoichiometry were as follows: with a longer chalcogenization time in 0.05 mol/L H2TeS4O6, the bandgap decreases from 2.39 eV to 1.70 eV, and from 2.15 eV to 1.52 eV in 0.1 mol/L H2TeS4O6, respectively (Figure 4, Table 5). The determined bandgap values were higher when compared to the bulk values (Table 4). This shows that mixed CdTe–CdS chalcogenides are formed. Tellurium as a metaloid has a very low bandgap value (approximately 0.33 eV) [40]. Even though diffractograms did not detect elemental tellurium, the molar ratio of Cd: Te: S in the layers shows there should be amorphous tellurium. This determines the decrease of the bandgap value of the layers when the concentration of the solution of the precursor and exposure time increases. At higher saturation of the PA 6 with cadmium and chalcogens species, a redshift of the absorption is observed in the spectra of the samples, which results in lower bandgap values. This is also due to the occurrence of an increased amount of the amorphous phase (Te) mixed with the nanocrystalline phase. Urbach energy changes from 0.56 eV to 0.62 eV with the increased treatment time in 0.05 mol/L H2TeS4O6 and from 0.58 eV to 0.66 eV in 0.1 mol/L H2TeS4O6, respectively (Figure 4, Table 5). The Urbach energy shows the density of states at the edge of the band and the local microstructural disorder. Therefore, the lower Urbach energy indicates fewer defects in structural bonds.
The mixed layers of cadmium chalcogenides should have more applications because its optical energy gap could be tuned by means of the composition between 1.52 and 2.36 eV.
From the FTIR spectra in the region of frequencies 500–1300 cm−1 of the polyamide surface before and after chalcogenization with monotelluropentathionic acid, it can be seen that the new material was formed or attached to the polyamide surface after chalcogenization (Figure 5). In the ranges of wave numbers 523–608 cm−1 and 1016–1204 cm−1, four peaks of the most intense bands are seen in the spectra of chalcogenized PA 6. S2O32− vibration can be assigned to six main absorption bands in the area of 520–1200 cm−1 [41]: for asymmetric deformation vibration, δas (O–S–O), for symmetric deformation vibration, δs(O–S–O), for symmetric valent, υs(S–O), and for asymmetric valent, υas(S–O).
Cadmium chalcogenide layers exhibited different IR spectra compared to an only-chalcogenized PA sample (Figure 5, curves S1-2 and S1-24, S2-2 and S2-24). As we can see, in the frequency region from 605 to 610 cm−1, the absorption of infrared light decreases, but it increases in the regions at 1013–1016 cm−1 and at 1203–1204 cm−1. Meanwhile, in the cases of increased PA 6 chalcogenization time and precursor concentration, the absorption of infrared light increased. Additionally, new peaks as shoulders in the region at 750–950 cm−1 appeared.
The broad band is shown with a maximum in the range 1209–1215 cm−1. The absorption at lower frequencies, with a peak at 609 cm−1, was strong enough, and the band was sharp and distinct so that it was combined with valence absorption. The band at 608–610 cm−1 can be assigned to the symmetric deformation O–S–O vibrations, δs(O–S–O), in the terminal SO3 groups of the polythionates [31].
The absorption peaks in the 790–1302 cm−1 region can be attributed to the asymmetric valence S–O vibrations, νas(S–O), and the symmetric valence S–O vibrations, νs(S–O), in the terminal SO3 groups of polythionates. [42].
However, the absorption intensity from 524 to 735 cm−1 increases significantly as chalcogenization proceeds. The bands with absorption peaks in this area correspond to symmetric deformation of O–S–O vibrations, δs(O–S–O), and asymmetric deformation vibrations of O–S–O, δas(O–S–O), in the terminal SO3 groups of polythionates [41].
The FTIR spectra propose only the probable course of the formation of thin chalcogen layers.
For the formation of nanolayers on the surface of PA, the main process is that the obtained layers of chalcogens having species, possibly chalcogenides, block the polymer surface, changing their properties. It can be assumed that new groups have formed, which are polychalcogenides and polychalcogenes. Further treatment with a cadmium salt solution leads to the formation of mixed CdTe–CdS layers, the FTIR spectra of which differ significantly from the spectra of chalcogenized PA 6 (Figure 5).
Some additional experiments are needed to determine the structural affiliation of the FTIR absorption bands. The transition metal chalcogenides have been extensively studied for their rich structural chemistry [43,44]. These compounds exhibit weak van der Waals interactions within a layer or chain, but a strong chalcogen-metal bond between these layers or chains. The chalcogen–metal bonds in transition metal chalcogenides are more covalent; therefore, chalcogen ions are less charged. The reduced electronegativity from S to Te is frequently accompanied by Q–Q bonding [45]. Different structural and physical properties could be related to the large ionic radius and reduced electronegativity.

3.3. XRD Characterization and the Mechanism of Cadmium Chalcogenides Layers Formation on the Substrate of Polyamide 6 Film

The structural characterization of the obtained CdTe–CdS layers was carried out using XRD. Due to the polycrystalline nature of the formed CdTe–CdS, as well as the simultaneous existence of many phases of cadmium telluride, cadmium sulfide, and sulfur of various chemical compositions and a high degree of crystallinity of the polyamide itself, complex diffraction patterns were obtained. The intensity of cadmium chalogenide peaks is several times lower than that of PA 6 peaks at 2θ < 28°. Thus, only the 30–70° region was used in detail in combination with JCPDS reference patterns and the available literature data [46,47,48,49].
Selected XRD diffractograms of cadmium chalcogenide layers on PA 6 film are shown in Figure 6, and the corresponding peak values are presented in Table 6. It can be seen that the cadmium chalcogenide layers consist of three phases: hexagonal cadmium telluride (CdTe, JCPDS card No 80-89), cubic cadmium telluride (CdTe, JCPDS card No 15-770), and hexagonal (CdS, JCPDS card No 80-6)) and monoclinic sulphur (S8, JCPDS card No 71-137). It can also be observed that as the chalcogenization time and H2TeS2O6 concentration increases, the amount of the mentioned phases in the cadmium chalcogenide layer also increases.
Because the details of the mechanism for both stages of treatment have not yet been elucidated, the proposed tentative mechanism is as follows.
At the first stage, TeS4O62− anions adsorb and diffuse into the surface of PA 6. In the second stage, chalcogenized samples reacted with cadmium sulfate solution and the formation of cadmium chalcogenide layer proceeds via [Cd(H2O)2(NH3)4]2+ ion reactions with the adsorbed TeS4O62− ion according to reactions (18)–(21):
2TeS4O62− + 5[Cd(H2O)2(NH3)4]2+ + 16OH → 2CdTe + 3CdS + 5SO42− + 18H2O + 20NH3
The fact that elemental sulfur was found in the cadmium chalcogenide layer can be explained by the fact that the anions of TeS4O62− in the surface of PA 6 are unstable at 80 °C. A part of them does not react with [Cd(H2O)2(NH3)4]2+ ions during the next step and decompose according to the reaction (19):
TeS4O62− + 4OH → 2S + 2SO42− + Te2− + 2H2O
The anions of SO42− wash out, and elemental sulfur remain in the polymer. Meanwhile, the anions of Te2− react with [Cd(H2O)2(NH3)4]2+ ions:
[Cd(H2O)2(NH3)4]2+ + Te2− → CdTe + 4NH3 + 2H2O
The total equation of the second step is:
3TeS4O62− + 6[Cd(H2O)2(NH3)4]2+ + 20OH → 3CdTe + 3CdS + 2S + 7SO42− + 22H2O + 24NH3
The procedures for attaching functionalized chelating polymers to the inner pore surface for the sorption of metal ions have been extensively studied [50,51]. A special role must also be attributed to an interaction of cadmium Cd2+ ions with the sorbed products from the chalcopolythionate anions. Sulfur or tellurium, either as free chalcogenide ions or when covalently bound to the amido group of PA, behave as a very effective Lewis base with respect to many transition metals, e.g., copper, iron, molybdenum, or zinc [52].

3.4. Morphological Characterization of Cadmium Chalcogenide Layers

SEM is a convenient technique for the investigation of cross-sections of PA 6 samples with cadmium chalcogenide layers. The SEM images for the cadmium chalcogenide layers on PA 6 show that the layers appear as homogeneous and dark strips, and their thickness depends on the time of the chalcogenization. The thickness of cadmium chalcogenide layers increases from 1.3 ± 0.1 μm to 1.7 ± 0.15 μm with the increase in the time of the chalcogenization from 2 to 24 h. The cross-sections (Figure 7) show that there is a good adherence of the layer to the polymer substrate.
The SEM micrograph shows a compact polycrystalline surface, consisting of one type of small, densely packed grains, unevenly distributed over a smooth homogeneous background. A small amount of compositional variation is observed in the micrographs of the deposited cadmium chalcogenide layers due to the longer exposure time in the precursor solution. A more detailed morphology was investigated by AFM, and the results are presented in Figure 8 and Table 7.
Figure 8 shows the top view 11.6 × 11.4 μm and 11.5 × 11.6 μm images of cadmium chalcogenides layers after 2 h (S2*-2) and 24 h (S2*-24) of chalcogenization, respectively. The surface images show (Figure 8) that the grains are unevenly distributed on the surface of the layers. Therefore, the interconnected grain particles merged together to form grains with an average diameter of ~30 nm. An agglomeration of grains takes place in the chalcogenide layers formed after a longer chalcogenization. The roughness parameters (Table 7) were determined using the section from the upper right position to the lower left of the top view image. RMS values of 29.97–35.21 nm at 279.2–293.0 nm maximum height allows determination of the layers of cadmium chalcogenides on PA 6 and demonstrate their sufficient high surface characteristics.
It is necessary to mention that mixed cadmium chalcogenide layers formed on PA 6 have specific compositional and morphological properties different from bulk chalcogenides, e.g., CdS or CdTe. The regularities established for the formation of those layers allow modulation in the formation of layers with desirable composition and morphology. Very simple procedures are required—the correct selection of single-source binary precursor, precursor concentration, and treatment time. The tunable bandgap is characteristic of these mixed binary layers of cadmium chalcogenides.
Such semiconductors of varying conductivity and bandgap values close to the solar spectrum are potential candidates for the fabrication of thin-film solar cells.

4. Conclusions

The successive ionic layer adsorption and reaction (SILAR) method was used to obtain ~1.3–1.7 μm cadmium telluride–cadmium sulfide layers on the surface of polyamide 6, and their optical, morphological, and crystalline phase information was determined. To obtain these layers, H2TeS2O6 0.05 and 0.1 mol/L solutions at 25 °C and alkali solution (pH 11.5) of 0.1 mol/L divalent cadmium in the complex ions form, [Cd(H2O)2(NH3)4]2+, at 80 °C were used. The quantity of Te, S, and Cd in the formed layers increases with increasing concentrations of chalcogens precursor solution and exposure time from 2 h to 48 h, respectively for Te ~0.672–0.934 µmol/cm2, for S ~1.3–2.1 µmol/cm2, and for Cd ~0.140–0.176 µmol/cm2. It was found that the initial rate of adsorption and the rate constants of adsorption and desorption increase with decreasing concentration of Te and Cd in the layers, except for S, where they decreased. The UV-Vis absorption spectra of the chalcogenides on PA 6 showed that the layers have high absorbance (≥104 cm−1), indicating direct bandgap transition. The optical study reveals that the direct transition bandgap for cadmium chalcogenide layers with different stoichiometry is found to be 1.70 to 2.36 eV for the samples chalcogenized in 0.05 mol/L solution of H2TeS4O6 and 1.52 to 2.15 eV for the samples chalcogenized in 0.1 mol/L solution. FTIR spectra of the modified polyamide show that, after chalcogenization, the new material is chemically bonded to the PA functional groups. XRD analysis of the cadmium chalcogenide layers on the PA 6 film showed that they consist of phases: hexagonal and cubic cadmium telluride, hexagonal cadmium sulfide, and monoclinic sulfur. The AFM and SEM studies showed that the mix layers of cadmium telluride and cadmium sulfide consist of crystalline grains with an average diameter of approximately 30 nm, some of which are agglomerated. Lower surface roughness parameters and a significantly smoother surface are observed because of shorter exposure time in the solution of the precursor. This correlates well with the calculated Urbach energy values, which show a lower microstructural disorder of the sample treated in the precursor solution for 2 h.

Author Contributions

Conceptualization, J.S. and L.S.; methodology, J.S., L.S. and R.I.; software, L.S. and R.I.; validation, J.S., L.S. and S.Z.; formal analysis, L.S. and R.I.; investigation, L.S.; resources, I.A.; data curation, J.S.; writing—original draft preparation, J.S. and L.S.; writing—review and editing, I.A. and R.I.; visualization, L.S. and S.Z.; supervision, J.S.; 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

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, X.; Wu, D.; Geng, H. Heterojunctions based on II-VI compound semiconductor one-dimensional nanostructures and their optoelectronic applications. Crystals 2017, 7, 307. [Google Scholar] [CrossRef] [Green Version]
  2. Ríos-González, J.A.; Mis-Fernández, R.; Camacho-Espinosa, E.; Riech, I.; Menéndez-Proupin, E.; Flores, M.A.; Orellana, W.; Peña, J.L. Inducing a level inside of CdTe bandgap doping with Sn using a co-sublimation technique by CSS. Mater. Sci. Semicond. Process. 2020, 107, 104836. [Google Scholar] [CrossRef]
  3. Halal, A.; Rahman, K.S.; Abdullah, S.F.; Sopian, K.; Amin, N. An investigation on CdS1-xTex interface compound in CdS/CdTe hetero-junction solar cells by density functional theory (DFT). Superlattices Microstruct. 2021, 151, 106805. [Google Scholar] [CrossRef]
  4. Supekar, A.; Kapadnis, R.; Bansode, S.; Bhujbal, P.; Kale, S.; Jadkar, S.; Pathan, H. Cadmium telluride/cadmium sulfide thin films solar cells: A review. ES Energy Environ. 2020, 10, 3–12. [Google Scholar] [CrossRef]
  5. Moreno-Regino, V.D.; Castañeda-de-la-Hoya, F.M.; Torres-Castanedo, C.G.; Márquez-Marín, J.; Castanedo-Pérez, R.; Torres-Delgado, G.; Zelaya-Ángel, O. Structural, optical, electrical and morphological properties of CdS films deposited by CBD varying the complexing agent concentration. Results Phys. 2019, 13, 102238. [Google Scholar] [CrossRef]
  6. Bosio, A.; Pasini, S.; Romeo, N. The history of photovoltaics with emphasis on CdTe solar cells and modules. Coatings 2020, 10, 344. [Google Scholar] [CrossRef] [Green Version]
  7. Nagaura, T.; Phan, H.P.; Malgras, V.; Pham, T.A.; Lim, H.; Ashok, A.; Kim, J.; You, J.; Nguyen, N.T.; Na, J.; et al. Universal electrochemical synthesis of mesoporous chalcogenide semiconductors: Mesoporous CdSe and CdTe thin films for optoelectronic applications. Angew. Chem. Int. Ed. 2021, 60, 9660–9665. [Google Scholar] [CrossRef]
  8. Maghouli, M.; Eshghi, H. Studying the effect of deposition time on physical properties of CdTe thin films; Influence of CdTe electrical properties on CdS/CdTe heterojunction rectifying behavior. Optik 2020, 218, 165132. [Google Scholar] [CrossRef]
  9. Wang, D.; Yang, Y.; Guo, T.; Xiong, X.; Xie, Y.; Li, K.; Li, B.; Ghali, M. Effect of pulse bias voltages on performance of CdTe thin film solar cells prepared by pulsed laser deposition. Sol. Energy 2021, 213, 118–125. [Google Scholar] [CrossRef]
  10. Saha, S.; Johnson, M.; Altayaran, F.; Wang, Y.; Wang, D.; Zhang, Q. Electrodeposition fabrication of chalcogenide thin films for photovoltaic applications. Electrochem 2020, 1, 19. [Google Scholar] [CrossRef]
  11. Maistruk, E.V.; Koziarskyi, I.P.; Koziarskyi, D.P.; Andrushchak, G.O. Optical properties of thin films cadmium chalcogenide obtained by the RF magnetron sputtering. Thirteen. Int. Conf. Correl. Opt. 2018, 10612, 277–284. [Google Scholar] [CrossRef]
  12. Carlos, V.-A. Self-powered cadmium chalcogenide photodetectors by pressurized air blast spraying. Opt. Mater. Express 2021, 11, 1104–1118. [Google Scholar] [CrossRef]
  13. Moger, S.N. Colour tunable co-evaporated CdSxSe1-x (0 ≤ x ≤ 1) ternary chalcogenide thin films for photodetector applications. Mater. Sci. Semicond. Process. 2020, 120, 105288. [Google Scholar] [CrossRef]
  14. Carcione, R.; Limosani, F.; Antolini, F. Cadmium telluride nanocomposite films formation from thermal decomposition of cadmium carboxylate precursor and their photoluminescence shift from green to red. Crystals 2021, 11, 253. [Google Scholar] [CrossRef]
  15. Yavorskyi, R.; Nykyruy, L.; Wisz, G.; Potera, P.; Adamiak, S.; Górny, S. Structural and optical properties of cadmium telluride obtained by physical vapor deposition technique. Appl. Nanosci. 2018, 9, 715–724. [Google Scholar] [CrossRef]
  16. Punitha, K.; Sivakumar, R.; Sanjeeviraja, C.; Sathe, V.; Ganesan, V. Physical properties of electron beam evaporated CdTe and CdTe:Cu thin films. J. Appl. Phys. 2014, 116, 213502. [Google Scholar] [CrossRef]
  17. Ivanauskas, R.; Milasiene, D. Fabrication of polyamide-Ag2Se composite films with controllable properties by an adsorption–diffusion method. J. Phys. Chem. Solids 2020, 145, 109533. [Google Scholar] [CrossRef]
  18. Ivanauskas, R.; Samardokas, L.; Mikolajunas, M.; Virzonis, D.; Baltrusaitis, J. Polyamide-thallium selenide composite materials via temperature and pH controlled adsorption-diffusion method. Appl. Surf. Sci. 2014, 317, 818–827. [Google Scholar] [CrossRef]
  19. Ancutiene, I.; Janickis, V.; Ivanauskas, R. Formation and characterization of conductive thin layers of copper sulfide (CuxS) on the surface of polyethylene and polyamide by the use of higher polythionic acids. Appl. Surf. Sci. 2006, 252, 4218–4225. [Google Scholar] [CrossRef]
  20. Šukytė, J.; Ivanauskas, R. Formation and properties of copper chalcogenides thin films on polymers formed using sodium telluropentathionate. Cent. Eur. J. Chem. 2013, 11, 1163–1171. [Google Scholar] [CrossRef] [Green Version]
  21. Ivanauskas, R.; Baltrusaitis, J. Synthesis and surface properties of polyamide-CuxSe composite thin films. Appl. Surf. Sci. 2013, 283, 360–366. [Google Scholar] [CrossRef]
  22. Phasook, N.; Kamoldirok, S.; Yindeesuk, W. Optical properties of Mn-doped CdS thin films grown by the SILAR method. J. Phys. Conf. Ser. 2018, 1144, 012009. [Google Scholar] [CrossRef]
  23. Chaudhari, K.B.; Gosavi, N.M.; Deshpande, N.G.; Gosavi, S.R. Chemical synthesis and characterization of CdSe thin films deposited by SILAR technique for optoelectronic applications. J. Sci. Adv. Mater. Devices 2016, 1, 476–481. [Google Scholar] [CrossRef]
  24. Hamid, Z.A.; Hassan, H.B.; Hassan, M.A.; Mourad, M.H.; Anwar, S. Deposition, characterization, performance of cadmium sulfide quantum dots thin films using SILAR technique for quantum dot sensitized solar cell applications. Key Eng. Mater. 2020, 835, 374–383. [Google Scholar] [CrossRef]
  25. Šukyte, V.J.; Ivanauskas, R.; Janickis, V. Preparation and some properties of Cu-Te-S thin films on the polyamide (PA) surface. Pol. J. Chem. 2005, 79, 759–771. [Google Scholar]
  26. Zalenkiene, S.; Sukyte, J.; Ivanauskas, R.; Janickis, V. Selenopentathionic and telluropentathionic acids as precursors for formation of semiconducting layers on the surface of polyamide. Int. J. Photoenergy 2007, 2007, 2660–2666. [Google Scholar] [CrossRef] [Green Version]
  27. Gregg, S.J.; Sing, K.S.W. Adsorption, Surface Area, and Porosity; Academic Press: London, UK, 1982; ISBN 0123009561. [Google Scholar]
  28. Olav Foss Salts of Monotelluropentathionic Acid. Available online: http://actachemscand.org/pdf/acta_vol_03_p0708-0716.pdf (accessed on 18 June 2021).
  29. Jeffery, G.H.; Passett, J.; Mendham, J.; Denney, R.C. Vogel’s Textbook of Quantitative Chemical Analysis. Available online: https://gtu.ge/Agro-Lib/Vogels_TEXTBOOK_OF_QUANTITATIVE_CHEMICAL_ANALYSIS_5th_ed__-_G_H_Jeffery.MsuCity.pdf (accessed on 18 June 2021).
  30. Ho, Y.S. Citation review of Lagergren kinetic rate equation on adsorption reactions. Scientometrics 2004, 59, 171–177. [Google Scholar]
  31. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process. Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  32. Purwanto, P.; Riska, R. Kinetics and equilibrium studies of electro adsorption of remazol red on modified stainless steel electrode. J. Phys. Conf. Ser. 2019, 1295, 012045. [Google Scholar] [CrossRef] [Green Version]
  33. Wu, F.C.; Tseng, R.L.; Juang, R.S. Characteristics of Elovich equation used for the analysis of adsorption kinetics in dye-chitosan systems. Chem. Eng. J. 2009, 150, 366–373. [Google Scholar] [CrossRef]
  34. Boyd, G.E.; Adamson, A.W.; Myers, L.S. The exchange adsorption of ions from aqueous solutions by organic zeolites. II. Kinetics. J. Am. Chem. Soc. 1947, 69, 2836–2848. [Google Scholar] [CrossRef]
  35. Alkan, M.; Demirbaş, Ö.; Doǧan, M. Adsorption kinetics and thermodynamics of an anionic dye onto sepiolite. Microporous Mesoporous Mater. 2007, 101, 388–396. [Google Scholar] [CrossRef]
  36. El Fawal, G.F.; Khalifa, R.E.; Rahman, S.A.; Eldin, M.S.M. Poly (Methacrylic acid) grafted regenerated cellulose ions exchangers membranes for Cu(II) ion adsorption: Kinetic, isotherm, and thermodynamic studies. Desalination Water Treat. 2020, 178, 182–192. [Google Scholar] [CrossRef]
  37. Hamdaoui, O.; Naffrechoux, E. Modeling of adsorption isotherms of phenol and chlorophenols onto granular activated carbon. Part I. Two-parameter models and equations allowing determination of thermodynamic parameters. J. Hazard. Mater. 2007, 147, 381–394. [Google Scholar] [CrossRef] [PubMed]
  38. Hassanien, A.S.; Akl, A.A. Effect of Se addition on optical and electrical properties of chalcogenide CdSSe thin films. Superlattices Microstruct. 2016, 89, 153–169. [Google Scholar] [CrossRef]
  39. Ghobadi, N. Band gap determination using absorption spectrum fitting procedure. Int. Nano Lett. 2013, 3, 1–4. [Google Scholar] [CrossRef] [Green Version]
  40. Strehlow, W.H.; Cook, E.L. Compilation of energy band gaps in elemental and binary compound semiconductors and insulators. J. Phys. Chem. Ref. Data 1973, 2, 163–200. [Google Scholar] [CrossRef] [Green Version]
  41. Foss, O. Structure of compounds containing chains of sulfur atoms. In Advances in Inorganic Chemistry and Radiochemistry; Emeleus, H.J., Sharpe, A.G., Eds.; Academic Press Inc.: New York, NY, USA, 1960; pp. 237–278. [Google Scholar]
  42. Nakamoto, K. Infrared and raman spectra of inorganic and coordination compounds. In Handbook of Vibrational Spectroscopy; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006. [Google Scholar]
  43. Zhang, W.; Wong, P.K.J.; Chua, R.; Wee, A.T.S. Growth and properties of magnetic two-dimensional transition-metal chalcogenides. In Spintronic 2D Materials: Fundamentals and Applications; Elsevier: Amsterdam, The Netherlands, 2019; pp. 227–251. ISBN 9780081021545. [Google Scholar]
  44. Jing, Y.; Liu, B.; Zhu, X.; Ouyang, F.; Sun, J.; Zhou, Y. Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications. Nanophotonics 2020, 9, 1675–1694. [Google Scholar] [CrossRef]
  45. Keane, P.M.; Frazen, H.F. Chalcogenides: Solid State Chemistry; King, R.B., Ed.; John Wiley, Sons: Chichester, UK; New York, NY, USA; Brisbane, Australia; Toronto, ON, Canada; Singapore, 1994; Volume 7. [Google Scholar]
  46. Nelmes, R.J.; McMahon, M.I.; Wright, N.G.; Allan, D.R. Observation of a high-pressure cinnabar phase in CdTe. Phys. Rev. B 1993, 48, 1314–1317. [Google Scholar] [CrossRef]
  47. Templeton, L.K.; Templeton, D.H.; Zalkin, A. Crystal structure of monoclinic sulfur. Inorg. Chem. 1976, 15, 1999–2001. [Google Scholar] [CrossRef] [Green Version]
  48. Swanson, H.E.; Morris, M.C.; Evans, E.; Ulmer, L. Cadmium telluride, CdTe (cubic). In Standard X-ray Diffraction Powder Patterns; National Bureau of Standards: Washington, DC, USA, 1964; Volume 25, p. 72. [Google Scholar]
  49. Yeh, C.Y.; Lu, Z.W.; Froyen, S.; Zunger, A. Zinc-blendewurtzite polytypism in semiconductors. Phys. Rev. B 1992, 46, 10086–10097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Zheng, C.; Wu, Q.; Hu, X.; Wang, Y.; Chen, Y.; Zhang, S.; Zheng, H. Adsorption behavior of heavy metal ions on a polymer-immobilized amphoteric biosorbent: Surface interaction assessment. J. Hazard. Mater. 2021, 403, 123801. [Google Scholar] [CrossRef]
  51. Mohammed, Y.A.; Ma, F.; Liu, L.; Zhang, C.; Dong, H.; Wang, Q.; Xu, X.; Al-Wahbi, A.A. Preparation of electrospun polyvinylidene fluoride/amidoximized polyacrylonitrile nanofibers for trace metal ions removal from contaminated water. J. Porous Mater. 2021, 28, 383–392. [Google Scholar] [CrossRef]
  52. Hestekin, J.A.; Bachas, L.G.; Bhattacharyya, D. Poly(amino acid)-functionalized cellulosic membranes: Metal sorption mechanisms and results. Ind. Eng. Chem. Res. 2001, 40, 2668–2678. [Google Scholar] [CrossRef]
Figure 1. Schematics of cadmium chalcogenide layer formation procedure.
Figure 1. Schematics of cadmium chalcogenide layer formation procedure.
Materials 15 00787 g001
Figure 2. Plots for the sorption of cadmium, tellurium, and sulfur having particles.
Figure 2. Plots for the sorption of cadmium, tellurium, and sulfur having particles.
Materials 15 00787 g002
Figure 3. UV-Vis absorption spectra of PA 6, chalcogenized samples in the solutions of H2TeS4O6 and samples with the layers of cadmium chalcogenides.
Figure 3. UV-Vis absorption spectra of PA 6, chalcogenized samples in the solutions of H2TeS4O6 and samples with the layers of cadmium chalcogenides.
Materials 15 00787 g003
Figure 4. Tauc plots (top) and Urbach plots (bottom) of cadmium chalcogenide layers on PA 6.
Figure 4. Tauc plots (top) and Urbach plots (bottom) of cadmium chalcogenide layers on PA 6.
Materials 15 00787 g004
Figure 5. FTIR absorption spectra of PA 6 chalcogenized samples in the solutions of H2TeS4O6 and samples with the layers of cadmium chalcogenides on the PA surface.
Figure 5. FTIR absorption spectra of PA 6 chalcogenized samples in the solutions of H2TeS4O6 and samples with the layers of cadmium chalcogenides on the PA surface.
Materials 15 00787 g005aMaterials 15 00787 g005b
Figure 6. X-ray diffraction patterns of the layers of cadmium chalcogenides on PA 6 films surface.
Figure 6. X-ray diffraction patterns of the layers of cadmium chalcogenides on PA 6 films surface.
Materials 15 00787 g006
Figure 7. SEM images of cross-sections of cadmium chalcogenide layers on PA 6.
Figure 7. SEM images of cross-sections of cadmium chalcogenide layers on PA 6.
Materials 15 00787 g007
Figure 8. AFM images of cadmium chalcogenide layers on PA 6.
Figure 8. AFM images of cadmium chalcogenide layers on PA 6.
Materials 15 00787 g008
Table 1. The main characteristics of Polyamide 6.
Table 1. The main characteristics of Polyamide 6.
PropertyDry/MoistUnit
Tensile strength at yield85/60MPa
Elongation at yield4%
Elongation at break70/200%
Hardness160/70N/mm2
Glass transition temperature60/5°C
Thermal conductivity (23 °C)0.23W/(K·m)
Density1.13g/cm3
Moisture absorption (23 °C)3%
Water absorption to equilibrium9.5%
Table 2. Duration of chalcogenization of PA 6 films at 25 °C and sample labelling.
Table 2. Duration of chalcogenization of PA 6 films at 25 °C and sample labelling.
Duration, hConcentration of H2TeS4O6, mol/L
0.050.1
Sample No
2S1-2S2-2
4S1-4S2-4
6S1-6S2-6
8S1-8S2-8
16S1-16S2-16
24S1-24S2-24
48S1-48S2-48
The sign “*” was added for the labeling of chalcogenized PA samples after their treatment with a cadmium salt solution (S1* and S2*).
Table 3. Model parameters for tellurium, sulfur, and cadmium species sorption kinetics by PA 6.
Table 3. Model parameters for tellurium, sulfur, and cadmium species sorption kinetics by PA 6.
ModelParametersTelluriumSulfurCadmium
H2TeS4O6 (mol/L)
0.050.100.050.100.050.10
Lagergrenr20.77860.90260.92650.90440.98430.8635
qe (μmol/cm2) 0.087
k1 (h−1) 0.261
APE (%) 46.7
Hor20.99610.99890.99960.99980.99990.9999
qe (μmol/cm2)0.6910.9931.3822.1480.1430.182
k2 (cm2·μmol−1·h−1)0.6180.2900.2180.4068.6023.312
ARE (%)8.13.11.72.61.61.7
Elovichr20.98250.9970.96220.96510.8530.9798
a (μmol·cm−2·h−1)11.6012.2471.491263.53 9.967
b (cm2·μmol−1)13.8507.0574.1985.133 57.471
APE (%)2.10.94.91.9 1.3
Boydr20.37570.7620.69270.73810.95950.7536
R (h−1) 0.289
APE (%) 11.6
Weber–Morrisr20.35370.60510.6210.16710.1220.4035
kint (μmol·cm−2·h−1/2)
APE (%)
Dumwald–Wagnerr20.71280.94580.96410.90450.99540.8996
K (h−1) 0.084 0.250
APE (%) 4.8 3.1
Table 4. Bandgap values of CdS and CdTe compounds [40].
Table 4. Bandgap values of CdS and CdTe compounds [40].
CompoundCdSCdTe
Bandgap, eV2.41–2.4251.5171.58
Temperature, K300300300
method of determinationphotoconductionelectroreflectionthermal activation
type of samplesingle crystallinesingle crystallinesingle crystalline
transitiondirect alloweddirectdirect allowed
Table 5. Optical properties of cadmium chalcogenide layers.
Table 5. Optical properties of cadmium chalcogenide layers.
Sample NoMolar Ratio Cd:Te:SDirect bandgap Eg (eV)R2Urbach Energy (eV)R2
S1*–21:4.3:5.42.360.99730.560.9993
S1*–241:4.3:8.61.700.99950.620.9998
S2*–21:5.2:13.02.150.99930.580.9997
S2*–241:5.1:12.21.520.99460.660.9994
Table 6. XRD peak assignments of cadmium chalcogenide layers on PA 6.
Table 6. XRD peak assignments of cadmium chalcogenide layers on PA 6.
Sample NoCdTeCdS
(JCPDS 80–6)
S8
(JCPDS 71–137)
S2*-2S2*-24(JCPDS 80–89)(JCPDS 15–770)
d (Å)d (Å)d (Å)h k ld (Å)h k ld (Å)h k ld (Å)h k l
35.6182.5192.5241 0 3
38.4442.34038.4442.340 2.4391 0 22.3411 4 2
39.0932.30239.0932.302 2.292 2 0 2.3041 2 4
40.7892.21040.7892.210 2.2133 2 3
45.3311.99945.3311.9991.9911 1 21.9543 1 1
60.7521.52360.7521.523 1.5131 0 4
Table 7. Surface roughness parameters of cadmium chalcogenide layers chalcogenized in 0.1 mol/L solution of H2TeS4O6 at 25 °C.
Table 7. Surface roughness parameters of cadmium chalcogenide layers chalcogenized in 0.1 mol/L solution of H2TeS4O6 at 25 °C.
Sample NoMax. Height, A (nm)Average Height, Zmean (nm)Average Roughness, Ra (nm)RMS Roughness, Rq (nm)Skewness, Rsk (nm)
Crude sample67.427.763.454.590.82
S2*-2279.262.0420.0929.971.26
S2*-24293.092.7225.8735.210.92
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ivanauskas, R.; Samardokas, L.; Sukyte, J.; Zalenkiene, S.; Ancutiene, I. Structural and Optical Characterizations of Cadmium Chalcogenide Layers on Polyamide Formed Using Monotelluropentathionic Acid. Materials 2022, 15, 787. https://doi.org/10.3390/ma15030787

AMA Style

Ivanauskas R, Samardokas L, Sukyte J, Zalenkiene S, Ancutiene I. Structural and Optical Characterizations of Cadmium Chalcogenide Layers on Polyamide Formed Using Monotelluropentathionic Acid. Materials. 2022; 15(3):787. https://doi.org/10.3390/ma15030787

Chicago/Turabian Style

Ivanauskas, Remigijus, Linas Samardokas, Judita Sukyte, Skirma Zalenkiene, and Ingrida Ancutiene. 2022. "Structural and Optical Characterizations of Cadmium Chalcogenide Layers on Polyamide Formed Using Monotelluropentathionic Acid" Materials 15, no. 3: 787. https://doi.org/10.3390/ma15030787

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