Skip to main content
ACS Polymers Au logoLink to ACS Polymers Au
. 2024 May 13;4(4):302–310. doi: 10.1021/acspolymersau.4c00018

Pattern Formation in Evaporating Polymer Solutions—Interplay between Dewetting and Decomposition

Pankaj Mahawar , M Praveena , Shreyanil Bhuyan , Dipin S Pillai , Sivasurender Chandran †,*
PMCID: PMC11328331  PMID: 39156558

Abstract

graphic file with name lg4c00018_0008.jpg

Pattern formation during solution evaporation is common in several industrial settings and involves a complex interplay of multiple processes, including wetting/dewetting, diffusion, and rheological characteristics of the solution. Monitoring the emergence of patterns during evaporation under controlled conditions may allow deconvolution of different processes and, in turn, improve our understanding of this common yet complex phenomenon. Here, we probe the importance of initial conditions, defined by the solution concentration c0, on the pattern formation in evaporating polymer solutions on the air–water interface. Intriguingly, the initial decrease in the lateral length scale (ξ), characterizing the patterns, takes an upturn at higher concentrations, revealing reentrant behavior. We employ a gradient dynamics model consisting of coupled evolution equations for the film height and the polymer fraction in the solution. Our simulations capture two different length scales revealing the reasons underlying the re-entrant behavior of ξ(c0). While the long-range destabilizing interactions between suspension and water result in the dewetting of thin film solutions, the phase separation between the polymer and solvent occurs at shorter length scales. Our results demonstrate the importance of initial concentration on pattern formation and, thereby, on the resultant properties of thin polymer films.

Keywords: pattern formation, dewetting, evaporation, Langmuir polymer films, gradient dynamics model, linear stability analysis, nonlinear dynamics simulations

Introduction

Pattern formation during the evaporation of suspensions is a common and recurring phenomenon. As an example, we often observe ring-like patterns formed by the evaporation of spilled-over suspensions like coffee, inks, or paints.13 Several industrial applications like large-scale coating,4,5 fabrication of solar cells,6 inkjet printing,7 disease detection,8 and forensic investigation of blood droplets9 rely on the understanding of such evaporation-induced patterns. Further, the evaporation of solutions containing one or more components provides an avenue to control pattern formation across several length scales.1019 Despite its enormous practical potential, the complex interplay of several physicochemical processes during evaporation limits our ability to fully harness evaporation to control pattern formation.11,2022

Evaporation is inherently a nonequilibrium process. The fraction of solutes in an initially dilute suspension progressively increases, resulting in variations in their flow behavior, which, in turn, could affect the wetting/dewetting rate and diffusion of solutes.4,11,20,21 Evaporation may create local concentration fluctuations, resulting in phase transitions like crystallization11,23,24 and glass transition.4,2527 In addition, the interaction of the suspension with the substrate dictates the contact line dynamics and thus reflects on the resulting patterns.28,29 To understand the development of evaporation-induced patterns, several theoretical approaches involving hydrodynamics and contact line dynamics,3032 gradient dynamics with lubrication approximation,3335 and moving boundary models were developed.3638 However, despite such progress, a complete understanding of the mechanisms underlying the pattern formation of evaporating solutions remains elusive.

Here, we probe the importance of initial conditions, defined by the solution concentration (c0), on pattern formation in evaporating polymer solutions. For a systematic control, we rely on a Langmuir trough, which allows the probing of the characteristics of the patterns before transferring the dried layers to a solid substrate. The dried films exhibit patterns such as flat islands with holes in between, networks, and domains, suggesting the dewetting of polymer solutions at the air–water interface. The initially constant film height (hm) increases monotonically with c0 beyond a critical concentration cc. Intriguingly, we observe that the lateral length scale (ξ), characterizing the patterns, displays a re-entrant behavior, i.e., the initial decrease in ξ takes an upturn at higher concentrations. We employ a gradient dynamics model consisting of coupled evolution equations for the height and polymer fraction in the suspension film. Our careful simulations, complemented by linear stability analysis, capture the growth of two different length scales in a narrow intermediate concentration range, while the low- and high-concentration suspensions exhibit the growth of a single length scale. While the long-range destabilizing interactions between suspension and water result in the dewetting of thin film solutions, the phase separation between the polymer and solvent occurs at shorter length scales. Nonlinear simulations revealed that decomposition, despite its lower growth rate, may nonlinearly dominate pattern formation at the intermediate concentrations. The change in ξ(c0) across the re-entrant behavior quantitatively agrees with our minimal model. Our results demonstrate that the initial concentration of the suspensions serves as a unique lever for quantitative control of the evaporation-induced patterns. Such explicit control over pattern formation would allow the tuning of various physical properties of polymer films.4,5,10

Experimental Section

Materials

We used thiol-terminated polystyrene with molecular weight Mw = 11 kg/mol and = 1.04, as purchased from Sigma-Aldrich. For dissolving polystyrene molecules, we used high-purity (≈99.8%) chloroform. We used deionized Milli-Q water with a resistivity of 18 MΩcm as a subphase in the Langmuir trough (Biolin Scientific). A Hamilton syringe with a capacity of 50 μL (purchased from Hamilton, Romania) was used to spread the polymer suspension onto the air–water interface. Polymer films as prepared were transferred onto a coverslip (received from Blue Star, India) sized at 22 × 22 mm2.

Methods

Substrate Preparation

Glass coverslips were cleaned using the RCA method to remove residual organic contamination and particles from the surface.39 Here, we briefly describe the procedure. A solution containing a 1:1:5 mixture of ammonium hydroxide (NH4OH), hydrogen peroxide (H2O2), and deionized water (H2O) was heated to a temperature of ca. 75 °C. Coverslips were immersed in this solution for 30 min and subsequently rinsed with plenty of deionized water. Such freshly treated surfaces were purged in dry nitrogen to ensure the complete evaporation of water.

Langmuir Technique

For all our experiments, we employed a Langmuir trough with a maximum surface area of 273 cm2. The temperature of the trough was maintained at ca. 25 °C. The entire system was enclosed within a closed chamber to prevent environmental contamination and minimize perturbation due to acoustic vibrations. Surface pressure is defined as Π = γ0 – γ, where γ0 stands for the surface tension of ultrapure Milli-Q water, and γ is the effective surface tension of the layer, which is proportional to the surface concentration of polymers. We utilized a platinum plate as the Wilhelmy plate for measuring Π. We used homogeneous solutions of polystyrene in chloroform with various concentrations ranging from 0.05 to 3 mg/mL. To ensure homogeneous dissolution, we stirred the solutions overnight. Chloroform facilitated the spreading of solution on the air–water interface due to its low boiling point and low surface tension. Using a Hamilton syringe, we spread the solutions drop-by-drop at various locations on the water surface. The total spreading volume was appropriately adjusted to maintain a constant total number of polymers for all of the experiments. The surface pressure was monitored closely while spreading. Newer drops were spread only after the fluctuations in surface pressure caused by the earlier drop completely relaxed to zero. We waited for at least 15 min after spreading to ensure a complete evaporation of the solvent. The floating polymer layers were compressed gently at a rate of 5 mm/min to avoid compression-induced instabilities and other artifacts at the surface. Surface pressure isotherms were recorded by plotting Π as a function of area A between the limiting barriers. To capture the patterns as they are formed, we transferred the films from the water surface to freshly cleaned glass coverslips using the Langmuir–Blodgett (LB) deposition method. To ensure optimal coverage and to avoid compression-induced pattern formation, we transferred all films at a minimal surface pressure Π = 10 mN/m.15 The substrates were lifted at a rate of 5 mm/min while simultaneously compressing at the same rate. This allowed us to minimize transfer-induced artifacts in the patterns. The deposited polymer films were completely dried by using a vacuum desiccator at ambient temperature.

Atomic Force Microscopy and Image Analysis

Surface morphology of the dried polymer films was probed by atomic force microscopy (AFM) in tapping (or semicontact) mode, using Origin MFP+3D (Oxford Instruments, UK). A cantilever with a resonance frequency of ca. 70 kHz, a force constant of 1.7 N/m, and a radius of curvature of ca. 7 nm was used for our measurements. We used a minimal normal force to avoid force-induced deformations. For a free oscillation voltage of 1 V, we typically used a set-point of ca. 0.7 V. The scanning rates were limited to 0.5 and 1 Hz to prevent scan instabilities and the dragging of the tip on the sample surface. Scanning was conducted at room temperature within a sealed chamber to mitigate contamination and to minimize acoustic vibrations. Image scanning was conducted at various locations across the polymer films, excluding the edges of the sample, to ensure that no artifacts resulting from LB transfer methods are inadvertently captured. The raw data were flattened using Gwyddion software.40 All images are presented as obtained without any additional processing. The resultant images are depicted with an appropriate color bar according to the height distribution of the patterns.

Experimental Results

Evaporation of Solutions with Different Initial Concentrations

The crucial aspect of evaporation-induced pattern formation is the spreading of polymer solutions at the air–water interface. The mobility of the underlying water surface avoids the pinning of the three-phase contact line of the droplets and hence allows homogeneous spreading on the surface. In our experiments, we systematically vary the initial solution concentrations c0 but fix the total number Npolymer of polymers by appropriately varying the volume of the solution spread on the water surface (see Figure 1a). The initial rate of spreading on the water surface results from a balance of the driving capillary and the resisting viscous forces,41,42 along with simultaneous evaporation. To understand this interplay between spreading and evaporation, it is instructive to compare the rates of both processes. For the initial dilute concentrations, the spreading rate is largely dominated by interactions between the solvent chloroform and the interface. Recent experiments reveal that chloroform evaporates at a rate of vev ≈ 0.1 μm/s.43 Using the values for the spreading parameter (|S| ≈ 10–2 N/m), contact angle (θ ≈ 0.1), and viscosity (η ≈ 10–3 Pa.s) from the literature,43 we obtain the initial spreading velocity,42vspread ∼ |S|·θ2/η, to be of the order of cm/s. Of course, any presence of convective flow of air at the air–water interface may increase the evaporation rate considerably. However, our experiments performed in an enclosed cabinet ensure negligible effects of airflow on evaporation. Therefore, in the initial stages of solution spreading, vevvspread suggests the domination of spreading over evaporation. Therefore, as captured in Figure 1b, we may conceive that the polymer solution undergoes an immediate transition from its initial drop-like shape to that of a flat thin film before any considerable evaporation can occur. Thereafter, with the evaporation of the solvent molecules, we expect a progressive increase in the concentration of polymers, and the spreading behavior will eventually be driven by the characteristics of the polymer, here polystyrene. The spreading parameter for polystyrene on water is negative.43,44 Thus, at later stages, we expect a complete dewetting of the thin film solution (Figure 1b). With an increase in the initial concentration, assuming a monotonic change in the spreading parameter, one may expect monotonic variation in the dewetting length scale. However, given that the number of molecules is held constant and that we have a mobile surface, do changes in the initial concentration reflect in pattern formation?

Figure 1.

Figure 1

Role of initial concentration on film characteristics: (a) the number (Npolymer) of polymers spread on the water surface is shown as a function of the initial concentration c0 with an appropriately adjusted volume. (b) Schematic representation of the key stages involved in the evaporation of the polystyrene solution on the water surface. (c) Surface pressure isotherms, Π vs A, for films formed from solutions of different c0 as defined in the panel in the units of mg/mL. The horizontal dashed line at Π = 10 mN/m highlights the pressure at which films are transferred. (d) Limiting area, A0, as a function of c0.

To answer this, we employ the water subphase of a Langmuir trough as the substrate for spreading polystyrene solutions at different initial concentrations. The Langmuir technique offers several advantages that make it an ideal experimental system for addressing this question: (a) the ability to gather the dewetted domains across the water surface together via the moving mechanical barriers, (b) the ability to characterize the nature of patterns and the evolution of mechanical interactions between them via surface pressure isotherms, and (c) the possibility of transferring the patterns onto solid substrates for further investigation. After spreading the solutions, we provided sufficient time to ensure complete evaporation of the solvent. Upon complete evaporation, dry dewetted polymer layers float on the surface. Such layers are compressed together to obtain surface pressure isotherms, as summarized in Figure 1c. All systems exhibit a range of area A, with no detectable surface pressure (Π ≈ 0), revealing an incomplete coverage of the dewetted patterns. With compression, i.e., on reducing the area between the barriers, Π increases monotonically, reflecting the interaction between the dewetted patterns. In line with earlier reports, we use a minimal compression rate (5 mm/min) to avoid compression-induced artifacts on the patterns.15 Interestingly, though Npolymer is constant for all solutions, the area A0 below which Π increases varies sensitively with the initial concentration c0 of the solution. We quantify this observation in Figure 1d by plotting A0 as a function of c0. There is an apparent decrease in A0 with an increase in c0, suggesting concentration-induced differences in the patterns floating on the water surface. We may attribute the decrease in A0 to the loosely packed patterns and/or the presence of thicker clusters with increasing c0. To understand this further, it is important to probe the patterns formed on the water surface. For this purpose, we transfer the patterns onto a solid surface at a minimal pressure Π = 10 mN/m to ensure that the compression has a negligible role in the self-assembled structures.15 The patterns as transferred are dried overnight in a vacuum to remove any traces of water. After appropriate drying, all films are investigated with an atomic force microscope.

Initial Concentration Affects Evaporation-Induced Patterns

We summarize the role of initial concentration in pattern formation in Figures 2, 3, and 4. As captured in Figure 2a, we observe flat structures with holes in between for c0 = 0.05 mg/mL. With an increase in c0 to 0.25 mg/mL, we notice an increased density of smaller holes (Figure 2b). On increasing c0 to 0.75 mg/mL, as captured in Figure 2c, we observe networks with underlying domain-like structures. With a further increase in c0 ≥ 1 mg/mL, domain-like patterns with an increasing lateral length scale and height are observed. At first glance, it is tempting to relate the concentration-induced differences in the patterns to the patterns frozen after different durations of dewetting.45 Dewetting of the film on an unfavorable surface is initiated by the rupture of the film to create holes that grow with time. Such growing holes may eventually coalesce to form networks. Upon small perturbations, the Rayleigh–Plateau instability drives the liquid filaments contained in the network to form domains, which may coalesce further to form larger domains. Thus, with time, we may expect a sequence of patterns: (i) holes in films, (ii) a network of filaments, (iii) the emergence of domain-like structures within the networks, and (iv) completely decoupled domains. While this sequence appears similar to the concentration-dependent transitions in patterns, there are important differences that need further attention. For instance, there is a rapid increase in the height of the patterns for concentrations c0 ≥ 0.75 mg/mL. In addition, the uniformity of the features in c0 = 0.25 mg/mL and c0 = 0.75 mg/mL suggest the presence of instabilities like spinodal dewetting.46,47 To further understand this behavior, we quantify the patterns via two prominent length scales: the mean height hm of the patterns and the lateral length scales ξ.

Figure 2.

Figure 2

Evaporation-induced pattern formation—role of initial concentration: representative surface morphologies of polystyrene films prepared from solutions with different initial concentrations: (a) c0 = 0.05 mg/mL, (b) c0 = 0.25 mg/mL, (c) c0 = 0.75 mg/mL, (d) c0 = 1.00 mg/mL, (e) c0 = 2.00 mg/mL, and (f) c0 = 3.00 mg/mL, respectively. Height profiles h(x) corresponding to the dashed lines in the images are shown below each micrograph. The dashed lines shown in h(x) vs x indicate the mean height (hm) of the patterns.

Figure 3.

Figure 3

Evolution of the mean height of the patterns with initial concentration: double logarithmic representation of the mean height hm (open symbols) of the patterns as a function c0. The solid red lines trace the concentration dependence of hm ≈ constant for c0 < cc and hmc0 for c0 > cc. The vertical dashed line marks the value of the critical concentration cc.

Figure 4.

Figure 4

Re-entrant transition in lateral length scales with initial concentration: (a) double logarithmic representation of the power spectral densities (PSDs) as a function of wave vector k. Data are shown for films formed from solutions of different c0 represented in the legend in units of mg/mL. The vertical dashed arrows point to the crossover wave vector, k = kc, capturing the transition from constant Guinier (k0) to Porod (k–4) behavior. (b) Double logarithmic representation of the largest lateral length scales ξ = 2π/kc as a function of c0. The dashed line is a guide to the eyes.

Let us begin with a discussion of the height of the patterns. Representative height profiles corresponding to all the images are shown in Figure 2. We observe a systematic increase in the height of the patterns with an increase in c0. As shown in Figure 3, we quantify this by plotting the mean height hm of the structures as a function of c0. Initially, for all concentrations c0 < cc = 0.6 mg/mL, the mean height hm ≈ 4 nm does not exhibit any dependence on concentration c0. We use polystyrene chains of molecular weight, Mw = 11 kDa, i.e., with the number of monomers N ≈ 106. Using Kuhn length of ak = 0.67 nm, we obtain the end-to-end distance of the chain Inline graphic6.9 nm and radius of gyration Inline graphic under melt conditions.48 Thus, hm ≈ 4 nm for c0 < cc reveals that the patterns formed are of polystyrene monolayers. For all initial concentrations c0 > cc, we observe hmc0. This is along the expected lines as the height of resulting layers ought to be proportional to the initial concentration.27,49

To characterize the lateral length scales defining the pattern, we rely on the power spectral density (PSD) analysis of the images. As captured in Figure 4a, we observe a transition from PSD ∼ k0 for k < kc to PSD ∼ k–4 for k > kc, where kc defines a crossover wave vector. This crossover from a constant PSD to a power law decrease is similar to that of the transition from Guinier to Porod behavior in the form factor considered in scattering.50,51 The wave vector independence of the PSD at low k is due to the finite size effects, and the power law decrease (PSD ∼ k–4) beyond a critical wave vector kc is because of the smooth surface of the scattering entities, here the patterns.50 Thus, the crossover wave vector allows us to capture the largest lateral correlation length scales, ξ = 2π/kc, defining the patterns. The crossover wave vector, kc, displays an intriguing dependence on the initial concentration. While kc for the lowest and the highest concentrations are similar, kc for intermediate concentrations is higher. This suggests a re-entrant transition in the lateral length scales with concentration. In Figure 4b, we plot ξ as a function of c0 for all the suspensions. For all concentrations, c0 < cc, ξ decreases systematically from ca. 300 nm down to ca. 70 nm. However, for c0 > cc, ξ takes an upturn and increases with an increase in c0 to a maximum of ca. 220 nm. We observe a clear re-entrant behavior of ξ with an increase in c0. How do we understand this re-entrant behavior of ξ(c0)?

Recent experiments on polymer-grafted nanoparticles revealed such re-entrant transition in lateral length scales.15 Supporting simulations based on the gradient dynamics model revealed that the bimodality corresponds to dewetting and concentration decomposition mediated by the interplay of enthalpic and entropic interactions.15 However, these simulations predicted that the length scales corresponding to these instabilities vary by more than 2 orders of magnitude.15,34 In our experiments, the maximum difference in length scales is only a factor of 3–4. While qualitatively, our experiments show similar reentrant behavior, the actual length scales differed by orders of magnitude. Thus, to understand the re-entrant behavior, we develop a minimal model inspired by the gradient dynamics model of Thiele and co-workers.15,33,34

Theoretical Modeling

Model Description

Several recent works have demonstrated the effectiveness of the gradient dynamics model in capturing the instabilities in thin films of solutions or suspensions.33,34,52 As established in the previous section, the initial film spreading occurs with negligible evaporation and results in a flat film at the air–water interface. The evolution of such a thin film can be described using gradient dynamics for the conserved fields, i.e., the film height h(x, t) and the local amount of solute ψ(x, t) = h(x, t)ϕ(x, t), where ϕ(x, t) is the volume fraction of the polymer. The effective free energy of the polymer solution thin film is thus a functional of the film height h(x, t) and the local amount of solute ψ(x, t). The free energy functional, F[h,ψ], is given by52

graphic file with name lg4c00018_m003.jpg 1

Here, γ is the surface tension of the solutions, and σ is the interfacial stiffness. The first term in eq 1 defines the free surface energy. The second term in eq 1 is the dewetting potential describing long-range destabilizing forces between the solution film and the substrate. This potential is defined as

graphic file with name lg4c00018_m004.jpg 2

where AH is the Hamaker constant and he is the equilibrium height that minimizes the dewetting potential. The third term in eq 1 signifies the bulk interaction potential capturing the possible concentration-dependent phase separation between polymer and solvent. Motivated by the Cahn–Hilliard theory of decomposition,53 we use a double well potential of the form ∼ (ϕ(ϕ – 1))2. Scaling this double well form with the energy density, we obtain

graphic file with name lg4c00018_m005.jpg 3

Here, kB is the Boltzmann constant, T is the temperature of the system, and a is the molecular size. The measure of the strength of the interaction, B, is tuned to match the experimental conditions. Unlike earlier works,15,34 we neglect the contribution of the conformational entropy of the chains in the interaction potential due to the short chain length present in our system. The last term in eq 1 defines the energetic penalty due to concentration gradients.

The temporal evolution of the coupled conserved fields, h(x, t) and ψ(x, t), is related to the spatial gradients of the variational derivative of the free energy functional F[h,ψ] as34

graphic file with name lg4c00018_m006.jpg 4
graphic file with name lg4c00018_m007.jpg 5

Here, Inline graphic and Inline graphic are the variational derivatives of energy functional F with respect to h and ψ, respectively. In the above, Qhh, Qhψ, Qψh, and Qψψ are the mobility factors and the element of the mobility matrix, Q, which is defined as follows

graphic file with name lg4c00018_m010.jpg 6

Linear Stability Analysis

To obtain insights into the resultant length scales, we rely on a linear stability analysis. This method evaluates the stability of a spatially uniform base state through modal analysis, wherein it is subjected to a periodic perturbation characterized by a wave vector, k. The system is considered unstable when it deviates from its base state upon perturbation, and the rate of this deviation is quantified by the growth rate (β). The wave vector (kmax) at which the growth rate reaches its maximum value determines the characteristic length scale of pattern formation.54 The wavelength, which is also the characteristic length scale, is computed as λmax (=2π/kmax). We obtain the solutions by perturbing the initially constant film height and volume fraction in the modal form as

graphic file with name lg4c00018_m011.jpg 7
graphic file with name lg4c00018_m012.jpg 8

The governing equation is nondimensionalized using appropriate scales, and the variables h and ψ are, respectively, perturbed about their base states and ψ̅ Inline graphic with extremely small amplitudes ϵ and ϵψ̃ as shown in eqs 7 and 8. Here, ϵ ≪ 1.

A set of two linear equations in and ψ̃ is obtained by linearizing the governing eqs 4 and 5 as follows

graphic file with name lg4c00018_m014.jpg 9

This system of linear equations can be converted to a 2 × 2 matrix eigenvalue problem with E11, E12, E21, and E22 as its elements. All these elements are functions of the wave vector (k). The eigenvalues of this matrix are computed to determine the growth rate (β). Owing to a two-dimensional matrix, two eigenvalues are obtained. For a given k, we obtain two real eigenvalues β, and the maximum of these is taken as the growth rate corresponding to k. The sign of β determines the stability of the system; a negative value implies that the system is stable, while a positive value implies that the system is unstable. From this analysis, we obtain the dispersion relation, i.e., the dependence of growth rate (β) on wave vector (k).

The values of the parameters used for the analysis are as follows: AH = 2 × 10–17 Nm, = 30 nm, ϕ̅ = 0.6, Inline graphic, a = 1 nm, γ = 27.15 mN/m, η = 0.0025 Pa·s, σ = aγ, B = 0.12, T = 300 K, and kB = 1.38 × 10–23 J/K. The vertical characteristic length scale l = 10 nm, the lateral length scale Inline graphic, and the time scale Inline graphic.

Theoretical Results: Linear Stability Analysis

The results of the linear stability analysis are summarized in Figure 5. The dispersion curves corresponding to different polymer fractions are shown in Figure 5a. At extreme polymer fractions, ϕ < 0.23 and ϕ > 0.77, we observe a single maximum in the dispersion curve suggesting the presence of a single mode of instability with wavelength λ1 = 2π/k1. However, for all intermediate polymer fractions 0.23 < ϕ < 0.77, we observe additional local maxima at higher wavevectors k2, suggesting the presence of another mode of instability with a smaller wavelength λ2 = 2π/k2. To highlight this observation, in Figure 5b, we display the concentration dependence of characteristic wavelengths λ1 and λ2. Thus, our linear stability analysis captures two different modes for intermediate polymer fractions. It is promising to note that the values of λ1 and λ2 are similar to the values of ξ shown in Figure 4b. However, the reasons underlying these instabilities need further discussion.

Figure 5.

Figure 5

Linear stability analysis capturing two modes of instabilities: (a) double logarithmic representation of the dispersion curves, growth rate β as a function of wavevector k, for solutions containing different volume fractions (ϕ) of polymer. While the lower and higher fractions of the polymer exhibit a single mode k1, intermediate fractions (red continuous lines) reveal an additional mode k2. (b) The concentration dependence of the corresponding length scales, λ1 = 2π/k1 and λ2 = 2π/k2, of these instabilities is shown.

Theoretical Results: Nonlinear Simulations

To understand the reason underlying the two modes of instability, we performed nonlinear simulations probing the spatiotemporal evolution of the two conserved variables h and ψ. As discussed earlier, we infinitesimally perturb the initially flat film, as given by eqs 7 and 8. For this simulation, we considered domain sizes that are integral multiples of the characteristic wavelengths, λ1 and λ2. To begin, we consider a domain size of 2λ1. No flow conditions are imposed at the boundaries, and periodic sinusoidal perturbations with a small amplitude (1%) are imposed as initial conditions. The finite difference method with 2000 regularly spaced grid points is employed to solve the differential equations, which was verified to be sufficient to obtain grid-independent solutions. The simulated spatiotemporal profiles of the film height (h) and the volume fraction of the polymer (ϕ) are shown in Figure 6. We observe a systematic evolution of h(x, t) (Figure 6a), while ϕ(x, t) is approximately constant (Figure 6b). This observation suggests that this corresponds to the dewetting instability of the thin film solution. The fact that the periodicity of h(x) matches exactly with λ1 reveals that λ1 captures the length scale corresponding to the dewetting mode. The results shown in Figure 6a,b capture solely the dewetting mode. How about the other instability corresponding to λ2? Because of the lower growth rate β corresponding to the second mode, we run the simulations for a longer time to allow for the manifestation of the mode corresponding to λ2. To reduce the computational cost, a domain size equal to λ2 is chosen. For this smaller domain size, 1000 grid points were chosen for the simulation. Clearly, for longer simulation runs, the polymer fraction evolves with time without much variations in h, suggesting an instability corresponding to concentration decomposition. The periodicity of ϕ(x), a shown in Figure 6c, is the same as that of λ2, revealing that the second mode of instability, observed at intermediate polymer fractions, corresponds to concentration decomposition.

Figure 6.

Figure 6

Nonlinear simulation capturing dewetting and decomposition: spatiotemporal evolution of (a) the nondimensional height profile, h(x, t), and (b) the volume fraction profile, ϕ(x, t), within a computational domain length of 2λ1. Different lines correspond to simulation runs for different times, as defined in panel (b). (c) Long-time simulations capturing the evolution of polymer fraction, ϕ, in a computational domain length of λ2 despite its low growth rate. For nondimensionalizing the variables, we use a lateral length scale, L = 16 nm, and a vertical length scale, l = 10 nm, and the time scale t0 = 6 ns.

Corroboration of Experiments and Simulations

As summarized in the previous section, simulations capture the coexistence of dewetting and decomposition instabilities for all intermediate polymer fractions. Interestingly, although the growth rates of decomposition modes are significantly smaller than that of the dewetting mode, our nonlinear simulations revealed the development of concentration decomposition for long waiting times. Thus, we define a single length scale λs* = λ2 for all intermediate fractions (0.23 < ϕ < 0.77) and λs* = λ1 for ϕ < 0.23 and ϕ > 0.77 capturing the dominant length scale in simulations. The concentration dependence of λs* is shown in Figure 7. Also shown is the experimental measure of the lateral length scales ξ as a function of the initial volume fraction (ϕ0) of polymers in the solution. At this point, it is important to highlight the differences between the experimental initial fraction of polymers ϕ0 and the fraction of polymers ϕ used in the simulations. Experimentally, ϕ0 characterizes the initial concentration of the solution. However, one can conceive that the dewetting or decomposition will initiate at later stages of evaporation, where the actual fraction of polymers would certainly be significantly higher than that of ϕ0. Unlike in experiments, the polymer fraction ϕ used in simulations is the fraction at which the instabilities occur. Thus, the differences observed between the experimental and theoretical intermediate fractions are expected. We may improve the situation by incorporating a sink term in the coupled evolution equations.55,56 The time dependence of the sink term could capture the effect of evaporation, which, in turn, may allow us to use the same definition for the concentration in both experiments and simulations. Despite the apparent differences, our simplified simulations impressively capture the experimental re-entrant behavior of ξ(ϕ0) by λs*(ϕ). Thus, a concentration-dependent transition from dewetting to decomposition and eventually to dewetting underlies the observation of the re-entrant behavior of ξ(ϕ0).

Figure 7.

Figure 7

Corroboration between experiments and simulations: quantitative agreement between experiments, ξ vs ϕ0, and simulations λs* vs ϕ. The extent of re-entrant in ξ is well captured by λs* with the fraction of polymers.

Conclusions

In conclusion, we have demonstrated the importance of initial conditions, defined via the initial concentration, on the length scale ξ characterizing the evaporation-induced patterns in thin polymer solutions. We reveal an intriguing concentration-dependent re-entrant behavior in the lateral length scales defining the patterns. Despite the simplifications of our model, we demonstrate that our simulation results are in quantitative agreement with the experimental results. Such a good corroboration allows us to reveal that the re-entrant behavior of ξ(ϕ0) is driven by the interplay of two instabilities: dewetting and decomposition. Our work provides an avenue to harness evaporation protocols to obtain hierarchical self-assembled patterns of polymers. Such differences in patterns would allow control over the resultant physical properties of polymer films.

Acknowledgments

We acknowledge the funding from the Startup Research Grant (SRG/2021/001276) and MATRICS grant (MTR/2022/000804), Science and Engineering Research Board. We acknowledge the Center for Nanosciences at IIT Kanpur for AFM measurements.

Author Contributions

§ P.M. and M.P. have equal contribution. CRediT: Pankaj Mahawar data curation, formal analysis, investigation, methodology, writing-original draft, writing-review & editing; Mullapudi Praveena data curation, formal analysis, investigation, methodology; Shreyanil Bhuyan formal analysis, methodology, validation, writing-original draft, writing-review & editing; Dipin S. Pillai formal analysis, funding acquisition, methodology, resources, supervision, writing-original draft, writing-review & editing; Sivasurender Chandran conceptualization, formal analysis, funding acquisition, project administration, resources, supervision, visualization, writing-original draft, writing-review & editing.

The authors declare no competing financial interest.

Special Issue

Published as part of ACS Polymers Auvirtual special issue “Polymer Science and Engineering in India”.

References

  1. Deegan R. D.; Bakajin O.; Dupont T. F.; Huber G.; Nagel S. R.; Witten T. A. Capillary flow as the cause of ring stains from dried liquid drops. Nature 1997, 389, 827–829. 10.1038/39827. [DOI] [Google Scholar]
  2. Thampi S. P.; Basavaraj M. G. Drying Drops of Colloidal Dispersions. Annu. Rev. Chem. Biomol. Eng. 2023, 14, 53–83. 10.1146/annurev-chembioeng-101121-085056. [DOI] [PubMed] [Google Scholar]
  3. Mampallil D.; Eral H. B. A review on suppression and utilization of the coffee-ring effect. Adv. Colloid Interface Sci. 2018, 252, 38–54. 10.1016/j.cis.2017.12.008. [DOI] [PubMed] [Google Scholar]
  4. Chandran S.; Baschnagel J.; Cangialosi D.; Fukao K.; Glynos E.; Janssen L. M.; Muller M.; Muthukumar M.; Steiner U.; Xu J.; et al. Processing Pathways Decide Polymer Properties at the Molecular Level. Macromolecules 2019, 52, 7146–7156. 10.1021/acs.macromol.9b01195. [DOI] [Google Scholar]
  5. Gu X.; Shaw L.; Gu K.; Toney M. F.; Bao Z. The meniscus-guided deposition of semiconducting polymers. Nat. Commun. 2018, 9, 534. 10.1038/s41467-018-02833-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Saki Z.; Byranvand M. M.; Taghavinia N.; Kedia M.; Saliba M. Solution-processed perovskite thin-films: the journey from lab-to large-scale solar cells. Energy Environ. Sci. 2021, 14, 5690–5722. 10.1039/D1EE02018H. [DOI] [Google Scholar]
  7. de Gans B.; Duineveld P. C.; Schubert U. S. Inkjet printing of polymers: state of the art and future developments. Adv. Mater. 2004, 16, 203–213. 10.1002/adma.200300385. [DOI] [Google Scholar]
  8. Rapis E. A change in the physical state of a nonequilibrium blood plasma protein film in patients with carcinoma. Technol. Phys. 2002, 47, 510–512. 10.1134/1.1470608. [DOI] [Google Scholar]
  9. Chen R.; Zhang L.; Zang D.; Shen W. Blood drop patterns: Formation and applications. Adv. Colloid Interface Sci. 2016, 231, 1–14. 10.1016/j.cis.2016.01.008. [DOI] [PubMed] [Google Scholar]
  10. Müller M.; Abetz V. Nonequilibrium processes in polymer membrane formation: Theory and experiment. Chem. Rev. 2021, 121, 14189–14231. 10.1021/acs.chemrev.1c00029. [DOI] [PubMed] [Google Scholar]
  11. Zhou J.; Man X.; Jiang Y.; Doi M. Structure Formation in Soft-Matter Solutions Induced by Solvent Evaporation. Adv. Mater. 2017, 29, 1703769. 10.1002/adma.201703769. [DOI] [PubMed] [Google Scholar]
  12. Rabani E.; Reichman D. R.; Geissler P. L.; Brus L. E. Drying-mediated self-assembly of nanoparticles. Nature 2003, 426, 271–274. 10.1038/nature02087. [DOI] [PubMed] [Google Scholar]
  13. Kinge S.; Crego-Calama M.; Reinhoudt D. N. Self-assembling nanoparticles at surfaces and interfaces. ChemPhysChem 2008, 9, 20–42. 10.1002/cphc.200700475. [DOI] [PubMed] [Google Scholar]
  14. Huang J.; Kim F.; Tao A. R.; Connor S.; Yang P. Spontaneous formation of nanoparticle stripe patterns through dewetting. Nat. Mater. 2005, 4, 896–900. 10.1038/nmat1517. [DOI] [PubMed] [Google Scholar]
  15. Sarika C.; Tomar G.; Basu J.; Thiele U. Bimodality and re-entrant behaviour in the hierarchical self-assembly of polymeric nanoparticles. Soft Matter 2015, 11, 8975–8980. 10.1039/c5sm02108a. [DOI] [PubMed] [Google Scholar]
  16. Bormashenko E.; Pogreb R.; Musin A.; Stanevsky O.; Bormashenko Y.; Whyman G.; Gendelman O.; Barkay Z. Self-assembly in evaporated polymer solutions: Influence of the solution concentration. J. Colloid Interface Sci. 2006, 297, 534–540. 10.1016/j.jcis.2005.11.025. [DOI] [PubMed] [Google Scholar]
  17. Bormashenko E.; Pogreb R.; Stanevsky O.; Bormashenko Y.; Stein T.; Gaisin V.-Z.; Cohen R.; Gendelman O. V. Mesoscopic patterning in thin polymer films formed under the fast dip-coating process. Macromol. Mater. Eng. 2005, 290, 114–121. 10.1002/mame.200400217. [DOI] [Google Scholar]
  18. Bormashenko E.; Pogreb R.; Stanevsky O.; Bormashenko Y.; Gendelman O. Formation of honeycomb patterns in evaporated polymer solutions: Influence of the molecular weight. Mater. Lett. 2005, 59, 3553–3557. 10.1016/j.matlet.2005.06.026. [DOI] [Google Scholar]
  19. Tönsmann M.; Scharfer P.; Schabel W. Critical solutal Marangoni number correlation for short-scale convective instabilities in drying poly (vinyl acetate)-methanol thin films. Polymers 2021, 13, 2955. 10.3390/polym13172955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Zang D.; Tarafdar S.; Tarasevich Y. Y.; Dutta Choudhury M.; Dutta T. Evaporation of a Droplet: From physics to applications. Phys. Rep. 2019, 804, 1–56. 10.1016/j.physrep.2019.01.008. [DOI] [Google Scholar]
  21. Parsa M.; Harmand S.; Sefiane K. Mechanisms of pattern formation from dried sessile drops. Adv. Colloid Interface Sci. 2018, 254, 22–47. 10.1016/j.cis.2018.03.007. [DOI] [PubMed] [Google Scholar]
  22. Routh A. F. Drying of thin colloidal films. Rep. Prog. Phys. 2013, 76, 046603. 10.1088/0034-4885/76/4/046603. [DOI] [PubMed] [Google Scholar]
  23. Park M. S.; Aiyar A.; Park J. O.; Reichmanis E.; Srinivasarao M. Solvent evaporation induced liquid crystalline phase in poly (3-hexylthiophene). J. Am. Chem. Soc. 2011, 133, 7244–7247. 10.1021/ja110060m. [DOI] [PubMed] [Google Scholar]
  24. Qian Q.; Yu S.; Li C. Y. Tuning Poly (l-lactic acid) Crystallization Pathways via Evaporative Crystallization on a Water Surface. Macromolecules 2023, 56, 8747–8753. 10.1021/acs.macromol.3c01769. [DOI] [Google Scholar]
  25. Peter S.; Meyer H.; Baschnagel J. Molecular dynamics simulations of concentrated polymer solutions in thin film geometry. II. Solvent evaporation near the glass transition. J. Chem. Phys. 2009, 131, 014903. 10.1063/1.3158607. [DOI] [PubMed] [Google Scholar]
  26. Rauch J.; Köhler W. Diffusion and thermal diffusion of semidilute to concentrated solutions of polystyrene in toluene in the vicinity of the glass transition. Phys. Rev. Lett. 2002, 88, 185901. 10.1103/PhysRevLett.88.185901. [DOI] [PubMed] [Google Scholar]
  27. Chandran S.; Handa R.; Kchaou M.; Al Akhrass S.; Semenov A. N.; Reiter G. Time allowed for equilibration quantifies the preparation induced nonequilibrium behavior of polymer films. ACS Macro Lett. 2017, 6, 1296–1300. 10.1021/acsmacrolett.7b00815. [DOI] [PubMed] [Google Scholar]
  28. Picknett R.; Bexon R. The evaporation of sessile or pendant drops in still air. J. Colloid Interface Sci. 1977, 61, 336–350. 10.1016/0021-9797(77)90396-4. [DOI] [Google Scholar]
  29. Sahoo S.; Mukherjee R. Pattern formation in evaporative drying of a polymer solution droplet over a soft swellable substrate. Langmuir 2023, 39, 9526–9537. 10.1021/acs.langmuir.3c01069. [DOI] [PubMed] [Google Scholar]
  30. Hu H.; Larson R. G. Evaporation of a sessile droplet on a substrate. J. Phys. Chem. B 2002, 106, 1334–1344. 10.1021/jp0118322. [DOI] [Google Scholar]
  31. Larson R. G. Instabilities in viscoelastic flows. Rheol. Acta 1992, 31, 213–263. 10.1007/BF00366504. [DOI] [Google Scholar]
  32. Hu H.; Larson R. G. Marangoni effect reverses coffee-ring depositions. J. Phys. Chem. B 2006, 110, 7090–7094. 10.1021/jp0609232. [DOI] [PubMed] [Google Scholar]
  33. Thiele U.; Todorova D. V.; Lopez H. Gradient dynamics description for films of mixtures and suspensions: Dewetting triggered by coupled film height and concentration fluctuations. Phys. Rev. Lett. 2013, 111, 117801. 10.1103/PhysRevLett.111.117801. [DOI] [PubMed] [Google Scholar]
  34. Sarika C.; Tomar G.; Basu J. Pattern formation in thin films of polymer solutions: Theory and simulations. J. Chem. Phys. 2016, 144, 024902. 10.1063/1.4939633. [DOI] [PubMed] [Google Scholar]
  35. Clarke N. Toward a model for pattern formation in ultrathin-film binary mixtures. Macromolecules 2005, 38, 6775–6778. 10.1021/ma0505777. [DOI] [Google Scholar]
  36. Harel M.; Taitelbaum H. Non-monotonic dynamics of thin film spreading. Eur. Phys. J. E 2021, 44, 69. 10.1140/epje/s10189-021-00017-w. [DOI] [PubMed] [Google Scholar]
  37. Okuzono T.; Ozawa K.; Doi M. Simple model of skin formation caused by solvent evaporation in polymer solutions. Phys. Rev. Lett. 2006, 97, 136103. 10.1103/PhysRevLett.97.136103. [DOI] [PubMed] [Google Scholar]
  38. Wodo O.; Ganapathysubramanian B. How do evaporating thin films evolve? Unravelling phase-separation mechanisms during solvent-based fabrication of polymer blends. Appl. Phys. Lett. 2014, 105, 153104. 10.1063/1.4898136. [DOI] [Google Scholar]
  39. Kern W. The evolution of silicon wafer cleaning technology. J. Electrochem. Soc. 1990, 137, 1887–1892. 10.1149/1.2086825. [DOI] [Google Scholar]
  40. Nečas D.; Klapetek P. Gwyddion: an open-source software for SPM data analysis. Open Phys. 2012, 10, 181–188. 10.2478/s11534-011-0096-2. [DOI] [Google Scholar]
  41. Mitchell J.; Weimer J. J. Controlled Spreading Rates to Distribute Nanoparticles as Uniform Langmuir Films. Langmuir 2021, 37, 5139–5150. 10.1021/acs.langmuir.1c00032. [DOI] [PubMed] [Google Scholar]
  42. Wyart F. B.; Martin P.; Redon C. Liquid/liquid dewetting. Langmuir 1993, 9, 3682–3690. 10.1021/la00036a053. [DOI] [Google Scholar]
  43. Fan X.; Wang C.; Fei L.; Zhou Q.; Xue Z.; Ma J.; Ma Y. Nascent Holes on Spin-Coated Polymer Nanofilms: Effect of Processing and Solvents. ACS Appl. Polym. Mater. 2022, 4, 2767–2782. 10.1021/acsapm.2c00070. [DOI] [Google Scholar]
  44. Intermolecular and Surface Forces, 3rd ed.; Israelachvili J. N., Ed.; Academic Press: San Diego, 2011; pp 661–674. [Google Scholar]
  45. Sharma A.; Reiter G. Instability of thin polymer films on coated substrates: rupture, dewetting, and drop formation. J. Colloid Interface Sci. 1996, 178, 383–399. 10.1006/jcis.1996.0133. [DOI] [Google Scholar]
  46. Seemann R.; Herminghaus S.; Jacobs K. Dewetting patterns and molecular forces: A reconciliation. Phys. Rev. Lett. 2001, 86, 5534–5537. 10.1103/PhysRevLett.86.5534. [DOI] [PubMed] [Google Scholar]
  47. Shiri R.; Schmeller L.; Peschka D.; Seemann R.; Wagner B. Impact of noise on spinodal dewetting of liquid-liquid films. Commun. Phys. 2023, 6, 109. 10.1038/s42005-023-01208-x. [DOI] [Google Scholar]
  48. Rubinstein M.; Colby R. H.. Polymer Physics; Oxford University Press, 2003. [Google Scholar]
  49. Meyerhofer D. Characteristics of resist films produced by spinning. J. Appl. Phys. 1978, 49, 3993–3997. 10.1063/1.325357. [DOI] [Google Scholar]
  50. Beaucage G. Approximations leading to a unified exponential/power-law approach to small-angle scattering. J. Appl. Crystallogr. 1995, 28, 717–728. 10.1107/S0021889895005292. [DOI] [Google Scholar]
  51. Hammouda B. Analysis of the Beaucage model. J. Appl. Crystallogr. 2010, 43, 1474–1478. 10.1107/S0021889810033856. [DOI] [Google Scholar]
  52. Diez J. A.; González A. G.; Garfinkel D. A.; Rack P. D.; McKeown J. T.; Kondic L. Simultaneous decomposition and dewetting of nanoscale alloys: A comparison of experiment and theory. Langmuir 2021, 37, 2575–2585. 10.1021/acs.langmuir.0c02964. [DOI] [PubMed] [Google Scholar]
  53. Cahn J. W.; Hilliard J. E. Free energy of a nonuniform system. I. Interfacial free energy. J. Chem. Phys. 1958, 28, 258–267. 10.1063/1.1744102. [DOI] [Google Scholar]
  54. Cross M.; Greenside H.. Pattern Formation and Dynamics in Nonequilibrium Systems; Cambridge University Press, 2009. [Google Scholar]
  55. Manzanarez H.; Mericq J.; Guenoun P.; Bouyer D. Modeling the interplay between solvent evaporation and phase separation dynamics during membrane. J. Membr. Sci. 2021, 620, 118941. 10.1016/j.memsci.2020.118941. [DOI] [Google Scholar]
  56. Thiele U. Thin film evolution equations from (evaporating) dewetting liquid layers to epitaxial growth. J. Condens. Matter Phys. 2010, 22, 084019. 10.1088/0953-8984/22/8/084019. [DOI] [PubMed] [Google Scholar]

Articles from ACS Polymers Au are provided here courtesy of American Chemical Society

RESOURCES