3.1. Deposition rates
Deposition rates of polymeric thin films were calculated from in situ film thicknesses measured by interferometry system during the polymerization. In the thickness estimation by interferometer, the refractive index of the polymer film is assumed to be constant and the change of the film thickness per fringe is calculated based on Frensel's equation and Snell's law (Eq. 1) [21].
\(\frac{\varvec{d}}{\varvec{f}\varvec{r}\varvec{i}\varvec{n}\varvec{g}\varvec{e}}=\frac{\varvec{\lambda }}{2{\varvec{\eta }}_{\varvec{p}\varvec{o}\varvec{l}\varvec{y}\varvec{m}\varvec{e}\varvec{r}}}\) (Eq. 1)
where, d, λ and η represent the estimated thickness value, the wavelength of the laser used in the interferometer system and the refractive index of the polymer, respectively. A laser with a wavelength of 633 nm was used and assuming the refractive index of the polymeric thin film to be 1.5, each d/fringe value in Eq. 1 was found to be equal to 211nm. After the depositions were completed, the accuracy of the measured interferometric thickness values was checked by an ex-situ profilometer. It was observed that the thickness values measured by both measurement methods were in perfect agreement. The deposition rates of thin films produced at different plasma power, substrate temperature and reactor pressure are given in Fig. 2.
At the same pressure and substrate temperature, when the plasma power increased from 20 to 60 W, the deposition rates of the films increased. This observation can be explained by Yasuda equation (Eq. 2) used to describe the plasma power per unit gas molecules in plasma polymerization [22].
MJ/Kg = (W/FM) × 1340 (Eq. 2)
where, W, F, and M represent the applied plasma power (watt), the flow rate of the monomer (sccm), and the molecular weight of the monomer (g/mol), respectively. According to Yasuda equation, as the applied plasma power under constant conditions increases, the energy input per molecule increases. Therefore, it is expected that increasing plasma power increases the deposition rate. However, as the plasma power increased from 60 to 80 W, a decrease in deposition rates was observed. It is known that in plasma polymerization studies, after a certain level of plasma power, a further increase in plasma power may cause ablation of the deposited films from the surface [23–25]. This could be the reason why the deposition rate decreased with increasing the plasma power from 60 to 80 W.
Reactor pressure is another important CVD parameter, which plays a significant role in the deposition rates [26, 27]. As can be seen in Fig. 2, higher deposition rates are obtained when the pressure increases from 100 to 300 mtorr. This can be attributed to the fact that as the pressure increases, the monomer vapor remains in the reactor for a longer time. The increase in the retention time may have increased the number of molecules remaining in the reactor per unit time and consequently the number of molecules adsorbed on the surface may have increased [28]. However, as the pressure was further increased to 500 mtorr, a decrease in the deposition rates was observed. This observed change in the deposition mechanism can be attributed to two possible reasons. The first one is that the further increase in the retention time of the monomer may have caused an increase in the rate of termination. The other possible reason is that the reduction in mean free path may have initiated some gas reactions [29]. The highest deposition rate was found to be 18.6 nm/min at a substrate temperature of 40°C, a reactor pressure of 300 mtorr, and a plasma power of 60 W. A typical PECVD polymerization process consists of several chemical and physical phenomena. First, monomer vapor is fed into the reactor and then gas diffusion takes place through the boundary layer. As a result of the contact of the gas with the substrate surface, deposition occurs on the substrate surface. Meanwhile, volatile components are removed from the surface. The slowest process step determines the deposition rate in the CVD process. The most known rate-limiting steps are adsorption (mass transfer) and surface reactions [30]. In this study, the deposition rate increased with increasing substrate temperature when all other PECVD parameters were kept constant. This result indicates that the rate-limiting step in PDPAEMA thin film synthesis is surface reactions.
In order to obtain more detailed findings on the deposition kinetic, the apparent activation energy of PDPAEMA thin film deposited at the highest deposition rate was calculated. In the CVD polymerization process, the conversion rate of monomer to polymer is very low [31, 32]. That is why, it can be assumed that the concentration of monomer in the reactor does not change. Since the monomer vapor concentration is too large to limit the deposition kinetics, the relationship between PECVD parameters and deposition rate can be simplified. Because the plasma power per molecule is also constant, it can be assumed that the deposition rate depends mainly on the substrate temperature. Based on this assumption, the deposition rate as a function of the substrate temperature can be plotted in an Arrhenius form (Eq. 3).
\({ln}k=-\frac{{E}_{a}}{RT}+ {ln}A\) (Eq. 3)
where, k is the reaction rate constant, Ea is the activation energy required for the reaction to occur (J/mol), R is the gas constant (8.314 J/Kmol), T is the reaction temperature (°K) and A is the frequency factor. Semilogarithmic graphic of deposition rates versus different substrate temperatures is presented in Fig. 3.
If the surface reaction rate is lower than the adsorption rate, the activation energy is positive. In the opposite case, the activation energy is negative. From Fig. 3, the activation energy of PDPAEMA was calculated as 17.56 kJ/mol. Since the reaction rate increases with increasing temperature, the activation value greater than zero is not surprising.
3.2. Film structures
The chemical structure of PDPAEMA thin film deposited at the highest deposition rate was revealed. The comparison of FTIR spectra of PDPAEMA thin film and DPAEMA monomer is given in Fig. 4. Both spectra were thickness-normalized and baseline-corrected. FTIR spectra of both PDPAEMA and DPAEMA monomer displayed the following major peak assignments: C-H stretching (3100 − 2800 cm− 1), C = O bond (1730 cm− 1), C-H bending (1500 − 1350 cm− 1), CH2 group (1260 cm− 1) [33, 34]. However, the C = C bond observed at 1650 cm− 1 in the monomer spectrum was not observed in the spectrum of the polymer. The absence of C = C bonds in the thin film indicates that the polymerization proceeded through C = C bonds without any entrained monomer in the as-produced polymer.
Looking at the FTIR spectrum of PDPAEMA, it is important to note here that, unlike PDPAEMA produced using different methods, some peak broadening and peak intensity were observed in this study. This is not surprising, because in PECVD polymerizations intense bombardments of ions, electrons, neutrals, etc. may lead to a certain degree of functional group loss and extensive fragmentation of the polymer [35, 36]. However, the large similarities in the main peaks between the monomer and the thin film imply that a high retention of the monomer structure. The chemical structure of the PDPAEMA thin film was also confirmed by XPS analysis. As expected, only C, O and N atoms were observed in the thin film in XPS survey scan analysis (Fig. 4b). The atomic percentages of C, O and N elements were found to be 75.7, 17.5 and 6.8 at.%, respectively. These values are very close to the values calculated in the chemical structure of DPAEMA monomer (79.2 at.% C, 13.2 at.% O, 6.6 at.% N). More detailed chemical investigation of PDPAEMA thin film was carried out by high-resolution mode of XPS. C1s spectrum of PDPAEMA can be curve-fitted into six components at binding energies of 289.1, 286.9, 286.1, 285.6, 285.1 and 284.7 eV, which can be attributed to -C*=O, -O-C*H2-, -N-C*H2-, -C*-(CH3)-, -C*H-N-, and -C*H3, respectively (Fig. 5a). O1s spectrum of PDPAEMA can be curve-fitted into two components at binding energies of 533.8 and 532.5 eV which can be attributed to -O* C*H3 and -C = O*, respectively (Fig. 5b) [37].
In order to investigate the response of PDPAEMA thin film to pH change, contact angles were measured after exposure to acidic and basic solutions. When PDPAEMA coated fabric was exposed to acidic solution, the fabric was completely wetted. After exposure to basic solution, PDPAEMA coated fabric showed hydrophobic properties, and the contact angle was measured as 114.7 °. This dramatic change in the contact angle measurements is attributed to the ionizable tertiary amine group in the PDMAEMA thin film, which is able to accept or donate proton according to the pH value of the solutions to which they are exposed [38]. From the studies in the literature, it is known that the pKa value of PDPAEMA thin film is in the range of 6.5 to 6.8 [39, 40]. In this study, when the pH value decreased to 3, tertiary amine groups were protonated and exhibited hydrophilic behavior due to the expansion of polymer chains with the effect of electrostatic repulsion force. When the pH value increased to 10, tertiary amine groups were deprotonated and PDPAEMA thin film exhibited hydrophobic behavior as a result of aggregation of the polymer chains. Depending on the ambient pH, it is very important that the change in the contact angle of pH responsive polymers is reversible. In order to display that the PDPAEMA thin film produced in this study has this capability, it was exposed to successively acidic and basic solutions. Figure 6a shows the contact angles of PDPAEMA coated fabric after successive exposure to acidic and basic solutions, similar contact angles were measured after each acid/base treatment cycle. Even after 5 repeated cycles, no significant change in the contact angle values was observed. Based on all these observations, it can be concluded that thin films are chemically stable and the conservation of functional groups in the monomer structure is high in the synthesized polymer. The predicted change in the chemical structure of the PDPAEMA thin film when exposed to acidic and basic solution is presented in Fig. 6b.
It is quite difficult to coat fragile and porous substrates such as fabric using traditional wet polymerization methods. In this study, PDPAEMA thin film was successfully coated on fabric by PECVD method due to its dry nature. SEM images of PDPAEMA coated and uncoated fabrics are shown in Fig. 7a-d. No difference was observed between the appearance of both fabrics. The porosity of the fabric is well preserved after coating, which indicates excellent conformal deposition.
PDPAEMA thin film produced by PECVD method can be used in areas where the use of it produced in bulk by conventional methods is very difficult or impossible. For example, PDPAEMA thin film produced at the nanoscale is expected to have high optical transmittance unlike the bulk polymer. Figure 8a shows photographs of uncoated and PDPAEMA thin film coated glass. No difference in the general appearance of the glasses was observed with the naked eye. The comparison of the optical transmittances of uncoated and PDPAEMA coated glasses is shown in Fig. 8b. It can be seen that PDPAEMA thin film does not cause any significant absorption and optical loss in the visible range.