Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-29T11:10:22.534Z Has data issue: false hasContentIssue false

Characterization of partial wetting by CMAS droplets using multiphase many-body dissipative particle dynamics and data-driven discovery based on PINNs

Published online by Cambridge University Press:  16 April 2024

Elham Kiyani
Affiliation:
Department of Mathematics, The University of Western Ontario, 1151 Richmond Street, London, ON, Canada N6A 5B7 The Centre for Advanced Materials and Biomaterials (CAMBR), The University of Western Ontario, 1151 Richmond Street, London, ON, Canada N6A 5B7
Mahdi Kooshkbaghi
Affiliation:
Simons Center for Quantitative Biology, Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, USA
Khemraj Shukla
Affiliation:
Division of Applied Mathematics, Brown University, 182 George Street, Providence, RI 02912, USA
Rahul Babu Koneru
Affiliation:
Department of Aerospace Engineering, University of Maryland, College Park, MD 20742, USA
Zhen Li
Affiliation:
Department of Mechanical Engineering, Clemson University, Clemson, SC 29634, USA
Luis Bravo
Affiliation:
DEVCOM Army Research Laboratory, Aberdeen Proving Ground, MD 21005, USA
Anindya Ghoshal
Affiliation:
DEVCOM Army Research Laboratory, Aberdeen Proving Ground, MD 21005, USA
George Em Karniadakis
Affiliation:
Division of Applied Mathematics, Brown University, 182 George Street, Providence, RI 02912, USA
Mikko Karttunen*
Affiliation:
The Centre for Advanced Materials and Biomaterials (CAMBR), The University of Western Ontario, 1151 Richmond Street, London, ON, Canada N6A 5B7 Department of Physics and Astronomy, The University of Western Ontario, 1151 Richmond Street, London, ON, Canada N6A 3K7 Department of Chemistry, The University of Western Ontario, 1151 Richmond Street, London, ON, Canada N6A 5B7
*
Email address for correspondence: mkarttu@uwo.ca

Abstract

The molten sand that is a mixture of calcia, magnesia, alumina and silicate, known as CMAS, is characterized by its high viscosity, density and surface tension. The unique properties of CMAS make it a challenging material to deal with in high-temperature applications, requiring innovative solutions and materials to prevent its buildup and damage to critical equipment. Here, we use multiphase many-body dissipative particle dynamics simulations to study the wetting dynamics of highly viscous molten CMAS droplets. The simulations are performed in three dimensions, with varying initial droplet sizes and equilibrium contact angles. We propose a parametric ordinary differential equation (ODE) that captures the spreading radius behaviour of the CMAS droplets. The ODE parameters are then identified based on the physics-informed neural network (PINN) framework. Subsequently, the closed-form dependency of parameter values found by the PINN on the initial radii and contact angles are given using symbolic regression. Finally, we employ Bayesian PINNs (B-PINNs) to assess and quantify the uncertainty associated with the discovered parameters. In brief, this study provides insight into spreading dynamics of CMAS droplets by fusing simple parametric ODE modelling and state-of-the-art machine-learning techniques.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (http://creativecommons.org/licenses/by-nc-sa/4.0), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is used to distribute the re-used or adapted article and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use.
Copyright
© The Author(s), 2024. Published by Cambridge University Press.

1. Introduction

Recent advancements in machine learning have opened the way for extracting governing equations directly from experimental (or other) data (Brunton, Proctor & Kutz Reference Brunton, Proctor and Kutz2016; Ren & Duan Reference Ren and Duan2020; Delahunt & Kutz Reference Delahunt and Kutz2022). One particularly exciting use of machine learning is the extraction of partial differential equations (PDEs) that describe the evolution and emergence of patterns or features (Lee et al. Reference Lee, Kooshkbaghi, Spiliotis, Siettos and Kevrekidis2020; Thiem et al. Reference Thiem, Kooshkbaghi, Bertalan, Laing and Kevrekidis2020; Meidani & Farimani Reference Meidani and Farimani2021; Kiyani et al. Reference Kiyani, Silber, Kooshkbaghi and Karttunen2022).

Spreading of liquids on solid surfaces is a classic problem (de Gennes Reference de Gennes1985; Bonn et al. Reference Bonn, Eggers, Indekeu, Meunier and Rolley2009). Although the theoretical foundations were laid by Young and Laplace already in the early 1800s (de Laplace Reference de Laplace1805; Young Reference Young1805), there are still many open questions, and it remains a highly active research field especially in the context of microfluidics (Nishimoto & Bhushan Reference Nishimoto and Bhushan2013) as well as in the design of propulsion materials (Jain et al. Reference Jain, Lemoine, Chaussonnet, Flatau, Bravo, Ghoshal, Walock and Murugan2021). As discussed in detail in the review of Popescu et al. (Reference Popescu, Oshanin, Dietrich and Cazabat2012), there are two fundamentally different cases: non-equilibrium spreading of the droplet, and the case of thermodynamic equilibrium when spreading has ceased and the system has reached its equilibrium state. In thermodynamic equilibrium, the Laplace equation relates the respective surface tensions of the three interfaces via (de Gennes Reference de Gennes1985; Popescu et al. Reference Popescu, Oshanin, Dietrich and Cazabat2012)

(1.1)\begin{equation} \cos{\theta_{eq}}= \frac{\gamma_{SG}-\gamma_{SL}}{\gamma_{LG}}, \end{equation}

where $\theta _{eq}$ is the equilibrium contact angle, and $\gamma _{SG}$, $\gamma _{SL}$ and $\gamma _{LG}$ are the surface tensions between solid–gas, solid–liquid and liquid–gas phases, respectively (see figure 1). Two limiting situations can be identified, namely, partial wetting and complete wetting. In the latter, the whole surface becomes covered by the fluid and $\theta _{eq} = 0^{\circ }$, that is, $\gamma _{SG}- \gamma _{LG} -\gamma _{SL} = 0$. When the equilibrium situation corresponds to partial wetting, $\theta _{eq} \ne 0^{\circ }$, it is possible to identify the cases of high-wetting ($0^{\circ } < \theta _{eq} < 90^{\circ }$), low-wetting ($90^{\circ } \le \theta _{eq} < 180^{\circ }$) and non-wetting ($\theta _{eq} = 180^{\circ }$).

Figure 1. A schematic showing the equilibrium contact angle (i.e. $\theta \equiv \theta _{eq}$), the surface tensions ($\gamma$), the threshold between low- and high-wetting regimes ($\theta _{eq} =90^{\circ }$), and a situation of a non-wetting droplet ($\theta _{eq} =180^{\circ }$). The last image demonstrates the occurrence of a precursor that is observed in some cases. In that case, the (macroscopic) contact angle is defined using the macroscopic part of the droplet, the cap in the rightmost image. The height of the precursor is in the molecular length scales (Hardy Reference Hardy1919; Nieminen et al. Reference Nieminen, Abraham, Karttunen and Kaski1992; Popescu et al. Reference Popescu, Oshanin, Dietrich and Cazabat2012).

When a droplet spreads, it is out of equilibrium, and properties such as viscosity and the associated processes need to be addressed (de Gennes Reference de Gennes1985; Bonn et al. Reference Bonn, Eggers, Indekeu, Meunier and Rolley2009; Popescu et al. Reference Popescu, Oshanin, Dietrich and Cazabat2012). In experiments, the most common choice is to use high-viscosity liquids in order to eliminate inertial effects. An early classic experiment by Dussan (Reference Dussan1979) gave a beautiful demonstration of some of the phenomena. She added tiny drops of marker dye on the surface of a spreading liquid, and observed a caterpillar-type rolling motion of the marker on the surface, giving rise to dissipation via viscous friction. Effects of viscosity and dissipation remain to be fully understood, and they have a major role in wetting phenomena (Cormier et al. Reference Cormier, McGraw, Salez, Raphaël and Dalnoki-Veress2012; McGraw et al. Reference McGraw, Chan, Maurer, Salez, Benzaquen, Raphaël, Brinkmann and Jacobs2016; Edwards et al. Reference Edwards, Ledesma-Aguilar, Newton, Brown and McHale2020).

A theoretical framework for comprehending and describing the behaviour of droplets during impact and spreading has been proposed by Gordillo, Riboux & Quintero (Reference Gordillo, Riboux and Quintero2019). They put forth a theoretical framework to elucidate the dynamics of spreading of Newtonian fluids. In this model, the spreading droplet is conceptualized as having a thick rim followed by a thin liquid film connected to the bulk of the drop. By employing principles of mass and momentum conservation along the rim, the liquid film and the droplet, a set of interconnected differential equations was formulated. The solution to these equations, considering appropriate boundary conditions, offers a thorough depiction of how the diameter of a Newtonian drops changes over time when impacting a smooth surface. Gorin et al. (Reference Gorin, Di Mauro, Bonn and Kellay2022) introduced a universal functional form to develop a model equation for estimating the time evolution of radial position, enabling predictions of transient spreading behaviour for both Newtonian and non-Newtonian fluids.

Calcium-magnesium-aluminosilicate (CMAS) is a molten mixture of several oxides, including calcia (CaO), magnesia (MgO), alumina (Al$_2$O$_3$) and silicate (SiO$_2$). It has a high melting point, typically around 1240 $^\circ$C (Poerschke & Levi Reference Poerschke and Levi2015) (although it can be significantly higher; see e.g. Wiesner, Vempati & Bansal (Reference Wiesner, Vempati and Bansal2016) and references therein), which allows it to exist in the molten state even at high temperatures encountered in modern aviation gas turbine engines (Clarke, Oechsner & Padture Reference Clarke, Oechsner and Padture2012; Ndamka, Wellman & Nicholls Reference Ndamka, Wellman and Nicholls2016). With high viscosity, high density and high surface tension, CMAS tends to form non-volatile droplets with $\theta _{eq}\ne 0$ rather than completely wetting the surfaces (Grant et al. Reference Grant, Krämer, Löfvander and Levi2007; Vidal-Setif et al. Reference Vidal-Setif, Chellah, Rio, Sanchez and Lavigne2012; Nieto et al. Reference Nieto, Agrawal, Bravo, Hofmeister-Mock, Pepi and Ghoshal2021). When it solidifies, it forms a glass-like material that can adhere to surfaces and resist erosion. The buildup of CMAS on turbine engine components can lead to clogging of the cooling passages and degradation of the protective coatings, resulting in engine performance issues and even damage or failure (Clarke et al. Reference Clarke, Oechsner and Padture2012; Ndamka et al. Reference Ndamka, Wellman and Nicholls2016; Song et al. Reference Song, Lavallée, Hess, Kueppers, Cimarelli and Dingwell2016; Wiesner et al. Reference Wiesner, Vempati and Bansal2016).

We simplified the CMAS wetting process by studying a fully molten CMAS droplet at a constant temperature, omitting considerations of surface chemical reactions, viscosity variations and phase changes. Existing analytical models that predict the spreading dynamics of highly viscous drops often rely on assumptions such as Stokes flow and completely wetting surfaces with zero contact angles. Our contribution lies in proposing a simple and general model for predicting the spreading dynamics for a broad range of systems, including partially wetting systems with non-zero contact angles.

The spreading of a droplet over a solid surface is commonly characterized using a power law $r \sim t^{\alpha }$, which expresses the radius of the wetted area as a function of time. The relationship is called Tanner's law for macroscopic completely wetting liquids at late times, with $\alpha = 1/10$ (Tanner Reference Tanner1979; Bonn et al. Reference Bonn, Eggers, Indekeu, Meunier and Rolley2009). Power laws have also been demonstrated at microscopic scales (Nieminen et al. Reference Nieminen, Abraham, Karttunen and Kaski1992). However, several conditions, such as surface properties, droplet shape and partial wetting, result in deviations from Tanner's law (McHale et al. Reference McHale, Shirtcliffe, Aqil, Perry and Newton2004; Cormier et al. Reference Cormier, McGraw, Salez, Raphaël and Dalnoki-Veress2012; Winkels et al. Reference Winkels, Weijs, Eddi and Snoeijer2012).

A common method, as the above suggests, for analysing the spreading dynamics is to investigate the existence of the power-law behaviour. To determine the presence of such power-law regimes in data, one can simply employ

(1.2)\begin{equation} \alpha (t) =\frac{{\rm d} \ln (r)}{{\rm d} \ln (t)} . \end{equation}

While this has worked remarkably well for complete wetting by viscous fluids, the situation for partial wetting is different (Winkels et al. Reference Winkels, Weijs, Eddi and Snoeijer2012). In our study, we investigate the spreading behaviour of CMAS droplets using multiphase many-body dissipative particle dynamics (mDPD) simulations. We generalize (1.2) such that it includes dependence on the initial droplet radius $R_0$ and $\theta _{eq}$ in order to describe partial wetting, that is, $\alpha \equiv \alpha (t, R_0, \theta _{eq})$.

Our objective is to gain a comprehensive understanding of the behaviour of CMAS droplet spreading dynamics by integrating knowledge about the fundamental physics of the system into the neural network architecture. To achieve this, we employ the framework of physics-informed neural networks (PINNs) (Raissi, Perdikaris & Karniadakis Reference Raissi, Perdikaris and Karniadakis2019), an emerging machine-learning technique that incorporates the physics of a system into deep learning. The PINNs address the challenge of accurate predictions in complex systems with varying initial and boundary conditions. By directly incorporating physics-based constraints into the loss function, PINNs enable the network to learn and satisfy the governing equations of the system.

The ability of PINNs to discover equations makes them promising for applications in scientific discovery, engineering design and data-driven modelling of complex physical systems (Karniadakis et al. Reference Karniadakis, Kevrekidis, Lu, Perdikaris, Wang and Yang2021). Their integration of physics-based constraints into the learning process enhances their capacity to generalize and capture the underlying physics accurately. Here, we also employ symbolic regression to generate a mathematical expression for each unknown parameter. Furthermore, we employ Bayesian physics-informed neural networks (B-PINNs) (Yang, Meng & Karniadakis Reference Yang, Meng and Karniadakis2021) to quantify the uncertainty of the predictions.

The rest of this paper is structured as follows. In § 2, we provide an overview of mDPD simulation parameters and system set-up. Simulation outcomes and the data preparation process are presented in § 3. Section 4 gives a brief introduction to the PINNs architecture, followed by a presentation of the results of PINNs and parameter discovery. The symbolic regression results and the mathematical formulas for the parameters are presented in § 5. Section 6 covers the discussion on B-PINNs as well as the quantification of uncertainty in predicting the parameters. Finally, we conclude with a summary of our work in § 7.

2. Multiphase many-body dissipative particle dynamics simulations

Three-dimensional simulations were performed using the mDPD method (Li et al. Reference Li, Hu, Wang, Ma and Zhou2013; Xia et al. Reference Xia, Goral, Huang, Miskovic, Meakin and Deo2017; Rao et al. Reference Rao, Xia, Li, McConnell, Sutherland and Li2021), which is an extension of the traditional dissipative particle dynamics (DPD) model (Español & Warren Reference Español and Warren1995; Groot Reference Groot2004); DPD is a mesoscale simulation technique for studies of complex fluids, particularly multiphase systems, such as emulsions, suspensions and polymer blends (Ghoufi & Malfreyt Reference Ghoufi and Malfreyt2012; Lei et al. Reference Lei, Bertevas, Khoo and Phan-Thien2018; Zhao et al. Reference Zhao, Chen, Zhang and Liu2021). The relation between DPD and other coarse-grained methods and atomistic simulations has been studied and discussed by Murtola et al. (Reference Murtola, Bunker, Vattulainen, Deserno and Karttunen2009), Li et al. (Reference Li, Bian, Yang and Karniadakis2016), Chan, Li & Wenzel (Reference Chan, Li and Wenzel2023) and Español & Warren (Reference Español and Warren2017).

In DPD and mDPD models, the position ($\boldsymbol {r}_i$) and velocity ($\boldsymbol {v}_i$) of a particle $i$ with mass $m_i$ are governed by Newton's equations of motion in the form

(2.1)\begin{equation} \left.\begin{gathered} \frac{{\rm d}\boldsymbol{r}_i}{{\rm d} t} = \boldsymbol{v}_i, \\ m_i\,\frac{{\rm d}\boldsymbol{v}_i}{{\rm d} t} = \boldsymbol{F}_{i} = \sum_{j\ne i}\boldsymbol{F}_{ij}^{C} + \boldsymbol{F}_{ij}^{D} + \boldsymbol{F}_{ij}^{R}. \end{gathered}\right\} \end{equation}

The total force on particle $i$, i.e. $\boldsymbol {F}_i$, consists of three pairwise components, i.e. the conservative $\boldsymbol {F}^{C}$, dissipative $\boldsymbol {F}^{D}$, and random $\boldsymbol {F}^{R}$ forces. The latter two are identical in DPD and mDPD models, given by

(2.2)$$\begin{gather} \boldsymbol{F}_{ij}^{D} =- \gamma\,\omega_{D}(r_{ij})\,( \boldsymbol{v}_{ij} \boldsymbol{\cdot} \boldsymbol{e}_{ij})\,\boldsymbol{e}_{ij}, \end{gather}$$
(2.3)$$\begin{gather}\boldsymbol{F}_{ij}^{R} = \zeta\,\omega_{R}(r_{ij})\,({\rm d}t)^{-1/2} \xi_{ij}\,\boldsymbol{e}_{ij}, \end{gather}$$

where $\boldsymbol {e}_{ij}$ is a unit vector, $\omega _{D}$ and $\omega _{R}$ are weight functions for the dissipative and random forces, and $\xi _{ij}$ is a pairwise conserved Gaussian random variable with zero mean and second moment $\langle \xi _{ij}(t)\,\xi _{kl} (t') \rangle = ( \delta _{ik} \delta _{jl} + \delta _{il} \delta _{jk})\,\delta (t-t')$, where $\delta _{ij}$ is the Kronecker delta, and $\delta (t-t')$ is the Dirac delta function. Together, the dissipative and random forces constitute a momentum-conserving Langevin-type thermostat. The weight functions and the constants $\gamma$ and $\zeta$ are related via fluctuation–dissipation relations first derived by Español & Warren (Reference Español and Warren1995),

(2.4)$$\begin{gather} \omega_{D} = (\omega_{R})^2, \end{gather}$$
(2.5)$$\begin{gather}\zeta = \sqrt{2\gamma k_{B}T}, \end{gather}$$

in which $k_{B}$ is the Boltzmann constant, and $T$ is the temperature. This relation guarantees the canonical distribution (Español & Warren Reference Español and Warren1995) for fluid systems in thermal equilibrium. The functional form of the weight function is not specified, but the most common choice (also used here) is

(2.6)\begin{equation} \omega_{D}(r_{ij}) = \left\{\begin{array}{@{}ll} (1 - r_{ij}/r_d)^s & \mbox{for}\ r_{ij} \le r_{d}, \\[2pt] 0 & \mbox{for}\ r_{ij}>r_{d}. \end{array}\right. \end{equation}

Here, $s=1.0$ is used, and $r_{d}$ defines a cutoff distance for the dissipative and random forces.

Although the above equations are the same for both DPD and mDPD, they differ in their conservative forces. Here, we use the form introduced by Warren (Reference Warren2001Reference Warren2003):

(2.7)\begin{equation} F^{C}_{ij} = A\,\omega_{C}(r_{ij})\,\boldsymbol{e}_{ij} + B (\rho_i + \rho_j) \omega_{B}\,\boldsymbol{e}_{ij}. \end{equation}

The functional form of both weight functions $\omega _C$ and $\omega _B$ is the same as $\omega _{D}$ in (2.6), but with different cutoff distances $r_c$ and $r_b$. The first term in (2.7) is the standard expression for the conservative force in DPD, and the second is the multi-body term. The constants $A$ and $B$ are chosen such that $A <0$ for attractive interactions, and $B>0$ for repulsive interactions; note that in conventional DPD, $A >0$ and $B=0$. The key component is the weighted local density

(2.8)\begin{equation} \rho_i = \sum_{j \ne i} \omega_\rho(r_{ij}). \end{equation}

There are several ways to choose the weight function (Zhao et al. Reference Zhao, Chen, Zhang and Liu2021), and here, the normalized Lucy kernel (Lucy Reference Lucy1977) in three dimensions is used:

(2.9)\begin{equation} \omega_\rho(r_{ij}) = \frac{105}{16 {\rm \pi}r^3_{{c}\rho}} \left(1+ \frac{3 r_{ij}}{r_{{c}\rho}}\right) \left(1- \frac{r_{ij}}{r_{{c}\rho}}\right)^3, \end{equation}

with a cutoff distance $r_{c\rho }$ beyond which the weight function $\omega _\rho$ becomes zero.

2.1. Simulation parameters and system set-up

To simulate molten CMAS, the parameter mapping of Koneru et al. (Reference Koneru, Flatau, Li, Bravo, Murugan, Ghoshal and Karniadakis2022) was used together with the open-source code LAMMPS (Thompson et al. Reference Thompson2022). In brief, the properties of molten CMAS at approximately 1260 $^\circ$C, based on the experimental data from Naraparaju et al. (Reference Naraparaju, Gomez Chavez, Niemeyer, Hess, Song, Dingwell, Lokachari, Ramana and Schulz2019), Bansal & Choi (Reference Bansal and Choi2014) and Wiesner et al. (Reference Wiesner, Vempati and Bansal2016), were used. In physical units, density was 2690 kg m$^{-3}$, surface tension 0.46 N m$^{-1}$, and viscosity 3.6 Pa s. Using the density and surface tension to estimate the capillary length ($\kappa = (\sigma /(\rho g))^{1/2}$) gives 4.18 mm. The droplets in the simulations (details below) had linear sizes shorter than the capillary length, hence gravity was omitted. In terms of physical units, time was $6.297\times 10^{-6}$ s, length was $17.017\times 10^{-6}$ m, and mass was $1.964 \times 10^{-8}$ kg.

Using the above values, droplets of initial radii $d = 8, 9 , 10, 11, 12$ in mDPD units, corresponding to $R_{0}= 0.136, 0.153, 0.17, 0.187, 0.204$ mm, respectively, were used in the simulations; all of them are smaller than the capillary length. The time step was 0.002 (mDPD units) corresponding to 12.59 ns. In addition, $k_{B}T =1$, $r_c =1.0$, $r_b=r_{{c}\rho }=0.75$, $r_d=1.45$, $\gamma =20$ and $B=25$. The attraction parameter $A$ in (2.7) has to be set for the interactions between the liquid particles ($A_{ll}$), and the liquid and solid particles ($A_{ls}$). The former was set to $A_{ll}=-40$, and $A_{ls}$ was chosen based on simulations that provided the desired $\theta _{eq}$, thus allowing for controlled variation of $\theta _{eq}$ (see figure 2).

Figure 2. The equilibrium contact angles $\theta _{eq}$ for the different attraction parameters between the liquid and solid particles ($A_{ls}$; see (2.7)). It is worth noting that the data for this figure have been extracted from Koneru et al. (Reference Koneru, Flatau, Li, Bravo, Murugan, Ghoshal and Karniadakis2022).

The initial configuration of the droplet and the solid wall was generated from a random distribution of equilibrated particles with number density $\rho =6.74$. This amounts to approximately 60 660 particles in the wall, and depending on the initial radius, anywhere between 14 456 and 48 786 particles in the droplet. Periodic boundary conditions are imposed along the lateral directions, and a fixed, non-periodic boundary condition is imposed along the wall-normal direction. Since mDPD is a particle-based method, the spreading radius and the dynamic contact angle are approximated using surface-fitting techniques. First, the outermost surface of the droplet is identified based on the local number density, i.e. particles with $\rho \in [0.45, 0.6]$. The liquid particles closest to the wall are fitted to a circle of radius $r$, i.e. the spreading radius. On the other hand, a sphere with the centroid of the droplet as the centre is fitted to the surface particles to compute the contact angle. The contact angle is defined as the angle between the tangent at the triple point (liquid–solid–gas interface) and the horizontal wall. The wall in these simulations is made up of randomly distributed particles to eliminate density and temperature fluctuations at the surface. Following Li et al. (Reference Li, Bian, Tang and Karniadakis2018), the root mean square height ($R_q$) of the surface scales linearly with $1/\sqrt {N_w}$, where $N_w=(2{\rm \pi} r_{cw}^3/3) \rho _w$ is the number of neighbouring particles. In this work, $R_q$ comes out to be around 0.0708 mDPD units, or $1.2\,\mathrm {\mu }$m.

As the CMAS droplet spreads on the substrate, it loses its initial spherical shape and begins to wet the surface as depicted in figure 3, forming a liquid film between the droplet and the substrate. Understanding how droplets behave on surfaces is important for a wide range of applications, including in industrial processes, microfluidics, propulsion materials, and the design of self-cleaning surfaces (Pitois & François Reference Pitois and François1999; Chen et al. Reference Chen, Bonaccurso, Deng and Zhang2016; Hassan et al. Reference Hassan, Yilbas, Al-Sharafi and Al-Qahtani2019; Jain et al. Reference Jain, Lemoine, Chaussonnet, Flatau, Bravo, Ghoshal, Walock and Murugan2021; Nieto et al. Reference Nieto, Agrawal, Bravo, Hofmeister-Mock, Pepi and Ghoshal2021).

Figure 3. (a) Illustration of the spreading behaviour of a CMAS droplet on a hydrophilic surface at different times. The droplet with initial size $R_0$ spreads on the surface with radius $r(t)$ and contact angle $\theta (t)$. (b) A series of snapshots from a simulation of a droplet with initial size $R_0 = 0.136$ mm and equilibrium contact angle $\theta _{eq} = 93.4^\circ$.

3. Simulation results

The size of a droplet changes over time. By tracking the changes, we can gain insight into the physical processes involved in spreading. The time evolution of the droplet radius ($r(t)$) is shown in figure 4. The log-log plots show the effect of the initial drop size $R_0$ and equilibrium contact angles $\theta _{eq}$ on the radius $r(t)$. Figure 4(a) displays $r(t)$ for initial drop sizes $R_0 = 0.136, 0.153, 0.17, 0.183, 0.204$ mm and equilibrium contact angle $\theta _{eq}=54.6^\circ$. Similarly, figure 4(b) shows the spreading radius for different equilibrium contact angles ($\theta _{eq}=93.4^\circ$, $85.6^\circ$, $77.9^\circ$, $70.1^\circ$, $62.4^\circ$, $54.6^\circ$, $45.3^\circ$ and $39.1^\circ$) with initial drop size $R_0=0.136$ mm.

Figure 4. The impact of $\theta _{eq}$ and $R_0$ on the droplet radii as a function of time for various (a) initial drop sizes with equilibrium contact angle $\theta _{eq} = 54.6^\circ$ corresponding to $A_{ls} = 30.0$, and (b) equilibrium contact angles (corresponding to $A_{ls} = -25.0, -25.8, -27.0, -28.0, -29.0, -30.0, -31.4, -32.2$) and initial drop size $R_{0}=0.136$ mm.

Eddi, Winkels & Snoeijer (Reference Eddi, Winkels and Snoeijer2013) used high-speed imaging with time resolution covering six decades to study the spreading of water–glycerine mixtures on glass surfaces. By varying the amount of glycerine, they were able to vary the viscosity over the range 0.0115–1.120 Pa s. They observed two regimes, the first one for early times with $\alpha$ changing continuously as a function of time from $\alpha \approx 0.8$ to $\alpha \approx 0.5$. This was followed by a sudden change to the second regime in which $\alpha$ settled to $0.1 < \alpha < 0.2$. As pointed out by Eddi et al. (Reference Eddi, Winkels and Snoeijer2013), the second regime agrees with Tanner's law (Tanner Reference Tanner1979). All of their systems displayed complete wetting.

Based on $r(t)$ of the CMAS drops and $\alpha$ shown in figure 4, it can be observed that $r(t)$ (and based on the power law, $\alpha$) depends on both the initial drop size and the equilibrium contact angle. The values of $\alpha$ for some simulation datasets are plotted over time in figure 5. The plot shows the behaviour of $\alpha$ for different initial drop sizes and equilibrium contact angles $\theta _{eq}=62.4^\circ$ and $85.6^\circ$.

Figure 5. The value of $\alpha$, calculated using (1.2), varies for different initial radii and fixed equilibrium contact angles (a) $\theta _{eq}=85.6^\circ$ and (b) $\theta _{eq}=62.4^\circ$. The figure illustrates that $\alpha$ is influenced by both the initial drop size $R_0$ and the equilibrium contact angle $\theta _{eq}$.

Inspired by the experimental results of Eddi et al. (Reference Eddi, Winkels and Snoeijer2013), the simulations of Koneru et al. (Reference Koneru, Flatau, Li, Bravo, Murugan, Ghoshal and Karniadakis2022) and the current simulations, we propose a simple sigmoid type dependence for $\alpha$:

(3.1) \begin{equation} \frac{{\rm d}\ln (r)}{{\rm d}\ln (t)} = \alpha(t, R_0, \theta_{eq}):= \eta \Big[\frac{1}{1+\exp(\beta (\tau-\ln(t)))}-1\Big]\!. \end{equation}

The two constant values of $\alpha$ discussed in the above references are the two extrema of the sigmoid curve, given that the transition between the two regimes occurs at $\ln (t_{transition})=\tau$.

The parameters of (3.1) are discovered by PINNs, and their dependence on $R_0$ and $\theta _{eq}$ is then expressed using symbolic regression. The general steps in the discovery of the droplet spreading equation and the extraction of the unknown parameters $\eta (R_0, \theta _{eq})$, $\beta (R_0, \theta _{eq})$ and $\tau ( R_0, \theta _{eq})$ are shown in figure 6 and can be summarized as follows.

  1. (i) Data collection: for this study, data are collected by conducting three-dimensional simulations using the mDPD method in LAMMPS with varying initial drop sizes $R_0$ and equilibrium contact angles $\theta _{eq}$.

  2. (ii) PINNs: the input of the network is time $t$, and the output of the network is the spreading radii $\tilde {r}(t)$. The physics-informed part of the PINNs is encapsulated in designing the loss function. In this study, the ‘goodness’ of the fit is measured by (a) deviation from trained data together with (b) deviation of network predictions and those from ordinary differential equation (ODE) (3.1) solutions. This optimization process reveals the the unknown parameters $\eta ( R_0, \theta _{eq})$, $\beta ( R_0, \theta _{eq})$ and $\tau ( R_0, \theta _{eq})$.

  3. (iii) Data interpolation: after PINNs are trained, we used their predictions together with two additional multilayer perceptron neural networks to fill the sparse parameter space. This step helps our next goal, which is relating the ODE parameters to $R_0$ and $\theta _{eq}$ without performing three-dimensional simulations.

  4. (iv) Symbolic regression: discovering a mathematical expression for each unknown parameter $\eta (R_0, \theta _{eq})$, $\beta (R_0, \theta _{eq})$ and $\tau ( R_0, \theta _{eq})$ of (3.1).

  5. (v) In order to quantify the uncertainty associated with our predictions, we utilize B-PINNs and leverage the insights gained from the PINNs prediction and symbolic regression, with a specific emphasis on the known value of $\eta$. By employing B-PINNs, we can effectively ascertain the values of two specific parameters, $\beta$ and $\tau$, which in turn enable us to quantify the uncertainty in our predictions.

Figure 6. The process of utilizing PINNs to extract three unknown parameters of the ODE (3.1), using three-dimensional mDPD simulation data. First, a neural network is trained using simulation data, where the input is time $t$ and the output is spreading radii $\tilde {r}(t)$. This neural network comprises four layers with three neurons, and is trained for 12 000 epochs. Subsequently, the predicted $\tilde {r}(t)$ is used to satisfy (3.1) in the physics-informed part. The loss function for this process consists of two parts: data matching and residual. By optimizing the loss function, the values of $\eta (R_0, \theta _{eq})$, $\beta (R_0, \theta _{eq})$ and $\tau (R_0,\theta _{eq})$ are determined for each set of $R_0$ and $\theta _{eq}$. After predicting the unknown parameters using PINNs, two additional neural networks, denoted as $NN_\beta$ and $NN_\tau$, are trained using these parameters to generate values for the unknown parameters at points where data are not available. The outputs of these networks, together with the outputs of the PINNs, are then fed through a symbolic regression model to discover a mathematical expression for the discovered parameter.

4. Physics-informed neural networks

In this section, we delve into the process of uncovering the parameters of the proposed ODE using the available simulation data.

In recent years, the concept of approximating the behaviours of complex nonlinear dynamical systems with experimental or numerically simulated data has gained significant traction, driven by the exponential growth in available data. The conventional approach for model fitting involves nonlinear least squares, which relies on iterative optimization algorithms (Srinivasan & Mason Reference Srinivasan and Mason1986). These algorithms can be computationally expensive and sensitive to initial conditions, sparsity, and high dimensionality of the data. To circumvent these issues, Brunton et al. (Reference Brunton, Proctor and Kutz2016) proposed a framework named SINDy (sparse identification of nonlinear dynamical system), which selects a class of algebraic functions along with an arbitrary order of derivatives. By performing sparse regression, which assumes that the contributions of many of these functions are negligible, SINDy can accurately discover ODEs and PDEs. On the other hand, equation learner (EQL; Martius & Lampert Reference Martius and Lampert2016) leverages the ability to use algebraic functions as neural network units, which brings interpretability to the model fitting process. Both SINDy and EQL share a reliance on predefined basis functions, constraining their ability to represent complex functional forms beyond those that are predefined. Therefore, applying these methods in the present study is not feasible. These coefficients not only are the coefficients of the basis functions but also cannot be utilized to define a library of functions, primarily due to the structure of the proposed ODE.

Physics-informed neural networks form another promising class of approaches that leverages the flexibility and scalability of deep neural networks to discover governing equations, incorporating physical laws (PDEs, ODEs, integro-differential equations (IDEs), etc.) and data (initial conditions plus boundary conditions) into the loss function (Raissi et al. Reference Raissi, Perdikaris and Karniadakis2019; Chen et al. Reference Chen, Lu, Karniadakis and Dal Negro2020; Mao, Jagtap & Karniadakis Reference Mao, Jagtap and Karniadakis2020; Mishra & Molinaro Reference Mishra and Molinaro2020; Shukla et al. Reference Shukla, Di Leoni, Blackshire, Sparkman and Karniadakis2020; Karniadakis et al. Reference Karniadakis, Kevrekidis, Lu, Perdikaris, Wang and Yang2021). In contrast, EQL and SINDy focus on deriving equations from data without enforcing the governing laws, lacking a guarantee of adherence to these principles. In addition, PINNs leverage automatic differentiation for computing the exact derivatives of differential equations, while SINDy, considered as an optimization method, relies on numerical differentiation for gradient approximations. The PINNs represent a more adaptable approach, capable of adjusting to complex systems by learning directly from data, even when a predefined mathematical model is completely or partially absent.

The method employed to find unknown parameters in this study exhibits potential for broader applications, such as different fluid properties, reduced-order PDEs, and scenarios in which only a part of the equation is absent (grey box learning framework; Kiyani et al. Reference Kiyani, Shukla, Karniadakis and Karttunen2023). While, potentially, the parameters of the proposed ODE could be discovered using other aforementioned approaches for the purpose of this paper, the scalability and extensibility to PDEs or more complex ODEs, and insensitivity to initial conditions in the fitting process compared to other methods, suggest that PINNs are the most suitable choice for this task.

4.1. Discovering parameters of the ODE

As discussed in § 3, our study aims to identify the values of the parameters $\eta (R_0,\theta _{eq})$, $\beta (R_0,\theta _{eq})$ and $\tau (R_0, \theta _{eq})$ in the ODE (3.1). The general steps of the framework are shown in figure 6(I). PINNs take time $t$ as an input to predict $\tilde {r}(t)$ at each time.

For each unique combination of a contact angle and initial drop size, the data were divided randomly into training, validation and test sets at ratio 80 : 10 : 10. A neural network with four layers of three neurons each was trained on the training data using the JAX and FLAX Python libraries on the M1 Apple chip. The training process consisted of 12 000 epochs, and took approximately 5 minutes to complete for each case.

The predicted values for the time evolution of the radii $\tilde {r}(t)$ should satisfy the data and the physics-informed step, i.e. meet the requirements of the ODE (3.1). The two-component loss function, designed to meet the requirements, consists of ${Loss_{data}}$ and ${Loss_{ODE}}$, given by

(4.1)\begin{equation} \left.\begin{gathered} {Loss_{data}}=\frac{1}{N_{k}} \sum_{i=1}^{i=N_{k}} | (r_{i}(t) - \tilde{r}_{i}(t)) | ^{2}, \\ {Loss_{ODE}} = \frac{1}{N_{r}} \sum_{i=1}^{i=N_{r}} \left | \frac{{\rm d}\ln (\tilde{r}) }{{\rm d}\ln (t)} - \eta \left(\frac{1}{1+\exp(-\beta(\ln(t)-\tau))}-1\right) \right | ^{2}, \end{gathered}\right\} \end{equation}

where $\tilde {r}(t)$ and $r(t)$ stand for the radii from the prediction and simulation, respectively. Here, $N_{k}$ is the number of training points, and $N_{r}$ is the number of residual points.

Figures 7, 8 and 9 illustrate the results obtained by utilizing PINNs to discover the parameters of the ODE (3.1), which describes the dynamics of the radii of the CMAS drops.

Figure 7. Comparison of the time evolution of the droplet radii: mDPD simulations (symbols), ODE model (3.1) (solid lines) and PINN predictions (dashed lines) for $\theta _{eq}=\{39.1^\circ, 62.4^\circ,93.4^\circ \}$ and $R_0=\{0.136, 0.153, 0.187,0.204\}$ mm parameter sets.

Figure 8. (a,b,c) The evolution of parameters $\eta (R_0,\theta _{eq})$, $\beta (R_0,\theta _{eq})$ and $\tau (R_0,\theta _{eq})$, respectively, over multiple epochs. These plots demonstrate that the parameters gradually converge to a stable state after $12\,000$ epochs. (d) The traces of the loss function for the PINNs framework. The learning curves demonstrate the decreasing trend of the loss functions, indicating that they converge to a stable point for all initial drop sizes and $\theta _{eq}$.

Figure 9. The values of (a) $\eta$, (b) $\beta$ and (c) $\tau$ obtained through PINNs. These values exhibit varying behaviours depending on the initial radius $R_{0}$ and equilibrium contact angle $\theta _{eq}$. The horizontal axes display the equilibrium contact angles $\theta _{eq}$. The vertical axes of all plots represent the values of $\eta$, $\beta$ and $\tau$. Here, $\eta$ remains nearly constant within a small range of values between $-0.325$ and $-0.200$, and $\beta$ and $\tau$ change within ranges from $1.0$ to $5.0$, and $6.5$ to $8.0$, respectively.

Figure 7 shows comparisons of the simulation data ($r(t)$), the prediction ($\tilde {r}(t)$) and the solution of the ODE (3.1). The figure demonstrates a remarkable degree of agreement between simulations, PINNs and our ODE model. The first three panels in figure 8 show the convergence of $\eta (R_0,\theta _{eq})$, $\beta (R_0,\theta _{eq})$, and $\tau (R_0,\theta _{eq})$ parameters during training. One can conclude that, all three parameters stabilize roughly after 12 000 epochs. In the rightmost panel of figure 8, the loss function (4.1) history is plotted against the training epochs, stabilizing around $10^{-4}$. This indicates successful training of the PINNs model. The results shown in figures 7 and 8 demonstrate the capability of our proposed framework to accurately predict the spreading radius of CMAS across the different initial radii and equilibrium contact angles.

Figures 9(a,b,c) show the values of $\eta (R_0,\theta _{eq})$, $\beta (R_0, \theta _{eq})$ and $\tau (R_0,\theta _{eq})$, respectively, obtained by PINNs for different initial radii $R_{0}$ ranging from $0.136$ to $0.204$ mm, and equilibrium contact angles $\theta _{eq}$ ranging from $39.1^\circ$ to $93.4^\circ$. The results show that $\eta$ changes within a small window between $-0.325$ and $-0.200$ for all $R_{0}$ and $\theta _{eq}$. However, the changes in $\beta$ (between $1$ and $5$) and $\eta$ (between $6.0$ and $8.0$) are significant, indicating that these parameters depend strongly on $R_{0}$ and $\theta _{eq}$.

4.2. Generate more samples of feasible radii and contact angles

As discussed earlier, the parameters in our ODE model (3.1) are functions of the initial radius and the equilibrium contact angle. Using PINNs, we were able to find the values for those parameters. To find a closed-form relation between the parameters $R_0$ and $\theta _{eq}$, more data than the rather small current set are needed. Performing three-dimensional mDPD simulations is, however, computationally expensive. In this subsection, we train two additional neural networks to capture the nonlinear relation between the ODE parameters found by PINNs, and the variables $R_0$ and $\theta _{eq}$. Then we will use these trained networks to fill our sparse parameter space to perform symbolic regression in the next section.

Specifically, we generate values for $R_{0}$ in the interval $[0.136, 0.204]\ \mathrm {mm}$ and $\theta _{eq}$ in the range $[40^\circ, 95^\circ ]$, as shown in figure 6(II).

Two fully connected networks $NN_{\beta }$ and $NN_{\tau }$ consist of eight dense layers with $256/256/256/128/64/32/16/8$ neurons. These networks are trained using an Adam optimizer with learning rate $10^{-2}$ for a total of 4000 epochs.

Discovered parameters from $NN_{\tau }$ and $NN_{\beta }$ are visualized in figures 10(a,b), respectively, where the parameter values obtained from PINNs are denoted by green and red circles, indicating the training data for $NN_{\beta }$ and $NN_{\tau }$, respectively. Furthermore, the parameter values generated by $NN_\beta$ and $NN_\tau$ are depicted as light orange and light blue dots. Visually, it is evident that these dots have filled the gaps between parameters that were absent in our LAMMPS dataset.

Figure 10. Predictions of the parameters (a) $\tau$ and (b) $\beta$ using the trained neural networks $NN_{\beta }$ and $NN_{\tau }$. The horizontal axes show $R_0$ and $\theta _{eq}$. The green and red circles correspond to the obtained values of $\tau$ and $\beta$ using the PINNs that were used to train $NN_{\beta }$ and $NN_{\tau }$. Additionally, the orange and blue dots represent the predicted values for grid interpolations from $R_0=0.136\ {\rm mm}$ to $R_0=0.204\ {\rm mm}$, and $\theta _{eq}=40^\circ$ to $\theta _{eq}=95^\circ$.

5. Symbolic regression

In this section, we use symbolic regression to find the explicit relation between the ODE parameters, and the initial radii and equilibrium contact angles. Symbolic regression is a technique used in empirical modelling to discover mathematical expressions or symbolic formulas that best fit a given dataset (Billard & Diday Reference Billard and Diday2002). The process of symbolic regression involves searching a space of mathematical expressions to find the equation that best fits the data. The search is typically guided by a fitness function that measures the goodness of fit between the equation and the data. The fitness function is optimized using various techniques, such as genetic algorithms, gradient descent, or other optimization algorithms. The equations discovered by symbolic regression are expressed in terms of familiar (i.e. more common) mathematical functions and variables, which can be understood and interpreted easily by humans.

In this study, we used the Python library gplearn for symbolic regression (Stephens Reference Stephens2016). As discussed in § 4.2, in order to have more accurate formulation for the ODE parameters before using symbolic regression, we trained two networks, $NN_{\beta }$ and $NN_{\tau }$, using the discovered values $\beta (R_0,\theta _{eq})$ and $\tau (R_0,\theta _{eq})$, enabling us to predict the parameter values for grid interpolations where no corresponding data points were available. The predicted parameters from both PINNs and the $NN_{\beta }$ and $NN_{\tau }$ networks are fed through the symbolic regression model to discover a mathematical formulation for each parameter. For this purpose, $\theta _{eq}$ and $R_0$ are fed as inputs, and $\eta$, $\beta$ and $\tau$ are the outputs. We set the population size to $5000$, and evolve $20$ generations until the error is close to $1\,\%$. Since the equation consists of basic operations such as addition, subtraction, multiplication and division, we do not require any custom functions.

The following results of symbolic regression can be substituted in (3.1):

(5.1)\begin{equation} \left.\begin{gathered} \eta =-0.255, \\ \beta = 0.283 + 0.27 \left(\frac{\theta_{eq}}{d}\right)\!, \\ \tau = 6.13 \left(\frac{d}{\theta_{eq}} + 1\right)\!, \end{gathered}\right\} \end{equation}

where $d$ is the initial size of the droplet in mDPD units.

Figure 11 shows the history of $\alpha (t, R_0, \theta _{eq})$ using the the ODE (3.1) with parameters from (5.1). The conversion between $d$ in mDPD units and $R_0$ in physical unit is $R_{0} = d\times 1.701\times 10^{-2}$ mm.

Figure 11. The behaviour of $\alpha$ from the right-hand side of (3.1) with parameters from (5.1): (a) different contact angles with fixed initial radius $R_0=0.136$ mm; (b) varying initial radii with fixed contact angle $\theta _{eq}=77.9^\circ$.

Figure 12(a) depicts the values of $\alpha$ for initial drop size $R_0 = 0.127$ mm and contact angles $\theta _{eq}=93.4^\circ$ and $87.2^\circ$. Figure 12(b)compares the solution of (3.1), using the discovered parameters (5.1), with the simulation data. The figure demonstrates the agreement between the ODE solution and the actual simulation results for this particular, unseen dataset. It is important to note that this particular drop size lies outside the training interval for initial drop sizes, $[0.136, 0.204]$ mm.

Figure 12. (a) The behaviour of the parameter $\alpha$ using (3.1) for $R_{0} = 0.127$ mm, which falls outside the range of the initial drop sizes used for training the networks. (b) The simulation data and the solution obtained from solving the ODE (3.1) with parameters from symbolic regression (5.1).

6. Bayesian physics-informed neural network results

Bayesian physics-informed neural networks integrate the traditional PINN framework with Bayesian neural networks (BNNs; Bykov et al. Reference Bykov, Höhne, Creosteanu, Müller, Klauschen, Nakajima and Kloft2021) to enable quantification of uncertainty in predictions (Yang et al. Reference Yang, Meng and Karniadakis2021). This framework combines the advantages of BNNs (Bishop Reference Bishop1997) and PINNs to address both forward and inverse nonlinear problems. By choosing a prior over the ODE and network parameters, and defining a likelihood function, one can find posterior distributions using Bayes’ theorem. The B-PINNs offer a robust approach for handling problems containing uncorrelated noise, and they provide aleatoric and epistemic uncertainty quantification on the parameters of neural networks and ODEs. The BNN component of the prior adopts Bayesian principles by assigning probability distributions to the weights and biases of the neural network. To account for noise in the data, we add noise to the likelihood function. By applying Bayes’ rule, we can estimate the posterior distribution of the model and the ODE parameters. This estimation process enables the propagation of uncertainty from the observed data to the predictions made by the model. We write (3.1) as

(6.1a)$$\begin{gather} \mathcal{N}_t(r; \boldsymbol{\lambda}) = f(t),\quad t \in \mathbb{R}^+, \end{gather}$$
(6.1b)$$\begin{gather}\mathcal{I} (r, \boldsymbol{\lambda}) = r_0, \quad t=0, \end{gather}$$

where $\boldsymbol {\lambda } = [\eta,\beta, \tau ]^\textrm {T}$ is a vector of the parameters of the ODE (3.1), $\mathcal {N}_t$ is a general differential operator, $f(t)$ is the forcing term, and $\mathcal {I}$ is the initial condition. This problem is an inverse problem; $\boldsymbol {\lambda }$ is inferred from the data with estimates on aleatoric and epistemic uncertainties. The likelihoods of simulation data and ODE parameters are given as

(6.2) \begin{equation} \left.\begin{gathered} P(\mathcal{D} \mid \boldsymbol{\theta}, \boldsymbol{\lambda}) = P(\mathcal{D}_{r} \mid \boldsymbol{\theta})\, P(\mathcal{D}_f \mid \boldsymbol{\theta}, \boldsymbol{\lambda})\, P(\mathcal{D}_{\mathcal{I}} \mid \boldsymbol{\theta}, \boldsymbol{\lambda}),\quad \text{where} \\ P(\mathcal{D}_{r} \mid \boldsymbol{\theta}, \boldsymbol{\lambda}) =\prod_{i=1}^{N_{r}} \frac{1}{\sqrt{2 {\rm \pi} \sigma_r^{(i)^2}}} \exp \Big[-\frac{({r}(\boldsymbol{t}_r^{(i)}; \boldsymbol{\theta}, \boldsymbol{\lambda})-\bar{r}^{(i)})^2}{2 \sigma_r^{(i)^2}}\Big]\!, \\ P(\mathcal{D}_f \mid \boldsymbol{\theta}, \boldsymbol{\lambda})=\prod_{i=1}^{N_f} \frac{1}{\sqrt{2 {\rm \pi} \sigma_f^{(i)^2}}} \exp \Big[-\frac{(\,f(\boldsymbol{t}_f^{(i)} ; \boldsymbol{\theta}, \boldsymbol{\lambda})-\bar{f}^{(i)})^2}{2 \sigma_f^{(i)^2}}\Big]\!, \\ P(\mathcal{D}_\mathcal{I} \mid \boldsymbol{\theta}, \boldsymbol{\lambda})= \prod_{i=1}^{N_\mathcal{I}} \frac{1}{\sqrt{2 {\rm \pi} \sigma_\mathcal{I}^{(i)^2}}} \exp \Big[-\frac{(\mathcal{I}(\boldsymbol{t}_i^{(i)} ; \boldsymbol{\theta}, \boldsymbol{\lambda})- \bar{\mathcal{I}}^{(i)})^2}{2 \sigma_{\mathcal{I}}^{(i)^2}}\Big]\!, \end{gathered}\right\} \end{equation}

where $D=D_r \cup D_f \cup D_{\mathcal {I}}$, with $\mathcal {D}_r=\{({\ln t}_r^{(i)}, {\ln r}^{(i)})\}_{i=1}^{N_r}$, $\mathcal {D}_f=\{(t_f^{(i)}, f^{(i)})\}_{i=1}^{N_f}$, $\mathcal {D}_\mathcal {I}= \{(t_\mathcal {I}^{(i)}, \mathcal {I}^{(i)})\}_{i=1}^{N_\mathcal {I}}$ are scattered noisy measurements. The joint posterior of $[\boldsymbol {\theta }, \boldsymbol {\lambda }]$ is given as

(6.3)\begin{align} P(\boldsymbol{\theta}, \boldsymbol{\lambda} \mid \mathcal{D}) &= \frac{P(\mathcal{D} \mid \boldsymbol{\theta}, \boldsymbol{\lambda})\, P(\boldsymbol{\theta}, \boldsymbol{\lambda})}{P(\mathcal{D})} \nonumber\\ &\approx P(\mathcal{D} \mid \boldsymbol{\theta}, \boldsymbol{\lambda})\, P(\boldsymbol{\theta}, \boldsymbol{\lambda}) \nonumber\\ &=P(\mathcal{D} \mid \boldsymbol{\theta}, \boldsymbol{\lambda})\, P(\boldsymbol{\theta})\,P(\boldsymbol{\lambda}). \end{align}

To sample the parameters from the the posterior probability distribution defined by (6.3), we utilized the Hamiltonian Monte Carlo approach (Radivojević & Akhmatskaya Reference Radivojević and Akhmatskaya2020), which is an efficient Markov Chain Monte Carlo (MCMC) method (Brooks Reference Brooks1998). For a detailed description of the method, see e.g. Neal (Reference Neal2011Reference Neal2012) and Graves (Reference Graves2011). To sample the posterior probability distribution, however, variational inference (Blei, Kucukelbir & McAuliffe Reference Blei, Kucukelbir and McAuliffe2017) could also be used. In variational inference, the posterior density of the unknown parameter vector is approximated by another parametrized density function, which is restricted to a smaller family of distributions (Yang et al. Reference Yang, Meng and Karniadakis2021). To compute the uncertainty in the ODE parameters by using B-PINN, a noise of $5\,\%$ was added to the original dataset. The noise was sampled from a normal distribution with mean $0$ and standard deviation ${\pm }1$.

Here, the neural network model architecture comprises of two hidden layers, each containing 50 neurons. The network takes time $t$ as the input, and generates a droplet radius $r(t)$ as the output. Additionally, we include a total of 2000 burn-in samples.

The computational expense of B-PINNs compared to traditional neural networks arises primarily from the iterative nature of Bayesian inference and the need to sample from the posterior distribution. The B-PINNs involve iterative Bayesian inference, where the posterior distribution is updated iteratively based on observed data. This iterative process requires multiple iterations to converge to a stable solution, leading to increased computational cost compared to non-iterative methods. Moreover, B-PINNs employ sampling-based algorithms such as MCMC or variational inference to estimate the posterior distribution of the model parameters. These algorithms generate multiple samples from the posterior distribution, which are used to approximate uncertainty and infer calibrated parameters.

Sampling from the posterior distribution can be computationally expensive, particularly for high-dimensional parameter spaces or complex physics models. Furthermore, B-PINNs often require running multiple forward simulations of the physics-based model for different parameter samples. Each simulation represents a potential configuration of the model parameters. Since physics-based simulations can be computationally intensive, conducting multiple simulations significantly increases the computational cost of training B-PINNs. Achieving a high acceptance rate for posterior samples, especially for high-dimensional data, demands running a large number of simulations. This further adds to the computational complexity.

Due to the computational expense associated with B-PINNs, we opt to use the method selectively for a few cases only. Utilizing the insights gained from PINNs prediction and symbolic regression, specifically the known value of $\eta = -0.255$, we can leverage the power of B-PINNs to uncover and ascertain the values of the parameters $\beta$ and $\tau$. Figure 13 showcases the comparison between the mean values of the radii $\tilde {r}(t)$ predicted by B-PINNs represented by solid lines, the corresponding standard deviations denoted by highlighted regions, and the simulation data used for training presented as stars, while the test data are indicated by coloured circles. The horizontal axes represent time, while the vertical axes depict the spreading radii $\tilde {r}(t)$. This comparative analysis is conducted for two distinct initial drop sizes, namely $R_0 = 0.136$ and $0.170$ mm, considering various equilibrium contact angles.

Figure 13. The mean and uncertainty (mean $\pm$2 standard deviations) of B-PINN predictions of the spreading radii history are given as solid lines and shaded regions, respectively. The test simulation data are depicted by solid circles, and training data are indicated by stars. This analysis is carried out for two different initial drop sizes, namely $R_0 = 0.136$ and $0.170$ mm, for three equilibrium contact angles.

Figure 14 illustrates the mean values of the parameters $\beta$ and $\tau$ obtained using B-PINNs along with their corresponding standard deviations. The solid lines represent the average values of the discovered parameters, while the highlighted regions indicate the standard deviations. The parameters discovered by PINNs are represented by the dashed lines. In figures 14(a,c), the vertical axes represent the values of $\beta$, while figures 14(b,d) display the values of $\tau$. The results are presented for two initial drop sizes, $R_0=0.136$ and $0.17$ mm. From the plots, it can be observed that the parameter $\beta$ exhibits a range of values between $1.0$ and $3.0$. On the other hand, the parameter $\tau$ fluctuates within the range $5.0$ to $7.0$. These ranges provide insight into the variability and uncertainty associated with the estimated values of $\beta$ and $\tau$ obtained through the B-PINN methodology.

Figure 14. Comparison between B-PINNs and PINNs discovered parameters for a range of equilibrium contact angles and two initial radii. The mean values (solid lines) and the standard deviations (mean values $\pm$2 standard deviations, shaded region) of (a,c) $\beta$ and (b,d) $\tau$. The dashed lines represent the parameters discovered by PINNs.

By comparing figures 9 and 14, it becomes evident that the discovered parameters $\beta$ and $\tau$ using PINNs of B-PINNs frameworks exhibit remarkable similarity. This striking similarity reinforces the efficacy and capability of our models in identifying accurately the parameters of the ODE (3.1). The close alignment between the discovered parameters in both figures demonstrates the robustness and reliability of our models. It highlights their ability to capture effectively the underlying dynamics and characteristics of the spreading behaviourof CMAS, leading to accurate parameter estimation. This consistency and agreement between the PINN and B-PINN results provide further validation of the power and effectiveness of our modelling approaches in uncovering the true values of the parameters $\beta$ and $\tau$ in the ODE. Additionally, (3.1) with anticipated parameters obtained using B-PINNs is solved using Odeint. The results are presented in figure 15, which provides a comparison between the simulated spreading radii $r(t)$ (circles) and the solution of the ODE (solid lines). This comparison is conducted for initial drop sizes $R_0 = 0.136$ and $0.17$ mm, considering different contact angles.

Figure 15. Comparison between the ODE solution with parameters found by B-PINNs (solid lines), and the simulation radii (circles). Two initial drop sizes, $R_0 = 0.136$ and $0.170$ mm, and three equilibrium contact angles are shown.

7. Conclusions

This study introduces a new approach to model the spreading dynamics of molten CMAS droplets. In the liquid state, CMAS is characterized by high viscosity, density and surface tension. The main objective is to achieve a comprehensive understanding of the spreading dynamics by integrating the underlying physics into the neural network architecture.

The study emphasizes the potential of PINNs in analysing complex systems with intricate dynamics. To study the dynamics of CMAS droplets, we performed simulations using the mDPD method. By analysing the simulation data and observing the droplet behaviour, we proposed a parametric equation (3.1), which consists of three unknown parameters. This parametric equation aims to capture and describe the observed behaviour of the CMAS droplets based on the simulation results. Using the data from the mDPD simulations, the study employed the PINNs framework to determine the parameters of the equation. Symbolic regression was then utilized to establish the relationship between the identified parameter values, and the initial droplet radii and contact angles. As a result, a simplified ODE model was developed, accurately capturing the spreading dynamics. The model's parameters were determined explicitly based on the droplet's geometry and surface properties. Furthermore, B-PINNs were employed to assess the uncertainty associated with the model predictions, providing a comprehensive analysis of the spreading behaviour of CMAS droplets.

In reality, the CMAS attack involves a complex interplay of reaction kinetics between CMAS and the thermal barrier coating, surface morphology (roughness, porosity) occurring under highly non-uniform thermal conditions. However, our findings extend beyond the isothermal conditions used in this study and even the specific case of CMAS droplets. The relationships uncovered and methods developed in this study have broader applications in understanding the spreading dynamics of droplets in general. By leveraging the insights gained from this research, one can investigate and understand the behaviour of droplets in diverse contexts, furthering our understanding of droplet spreading phenomena. Potentially, this knowledge can be used in developing strategies for effective droplet management and optimizing processes involving droplets in a wide range of practical applications.

The key aim of this work is to create a mathematical model for the partial wetting induced by CMAS droplets with specific properties. It is clear that this model has the capacity for expansion to encompass diverse fluid properties. By exploring different fluids with unique characteristics, we can employ the same methodology to uncover the ODE's correlation with fluid properties. For future work, we can extend the proposed equations to incorporate fluid properties by considering different fluids with different properties, as the same method can be used to find the dependency of the ODE to fluid properties. Also, it is worth pursuing a systematic comparison of our results with Gordillo's theory (Gordillo et al. Reference Gordillo, Riboux and Quintero2019; Gorin et al. Reference Gorin, Di Mauro, Bonn and Kellay2022). Additionally, in the case of a contact angle $\theta$ changing with time, a similar scaling law rationale can be applied, i.e. $\theta (t) \sim t^{\gamma }$ (see De Ruijter, De Coninck & Oshanin Reference De Ruijter, De Coninck and Oshanin1999). One can extend the current study to a system of ODEs that incorporates temporal variations in both radii and the advancing contact angle.

Funding

E.K. thanks the Mitacs Globalink Research Award Abroad and Western University's Science International Engagement Fund Award. M.Ka. thanks the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Canada Research Chairs Program. The work was partially supported by a US Department of Energy (DOE) grant (DE-SC0023389). R.B.K. would like to acknowledge the support received from the US Army Research Office Mathematical Sciences Division for this research through grant no. W911NF-17-S-0002. L.B. and A.G. were supported by the US Army Research Laboratory 6.1 basic research program in propulsion sciences. Z.L. and G.K. acknowledge support from the AIM for Composites, an Energy Frontier Research Center funded by the DOE, Office of Science, Basic Energy Sciences (BES) under Award no. DE-SC0023389. Z.L. also acknowledges support from the National Science Foundation (grants OAC-2103967 and CDS&E-2204011). Computing facilities were provided by the Digital Research Alliance of Canada (https://alliancecan.ca) and the Center for Computation and Visualization, Brown University. The authors acknowledge the resources and support provided by Department of Defense Supercomputing Resource Center (DSRC) through use of ‘Narwhal’ as part of the 2022 Frontier Project, Large-Scale Integrated Simulations of Transient Aerothermodynamics in Gas Turbine Engines. Work at Brown University and Clemson University was supported by AIM for Composites, an Energy Frontier Research Center funded by the DOE, Office of Science, BES under Award no. DE-SC0023389 (AI Models).

Declaration of interests

The authors report no conflict of interest.

References

Bansal, N.P. & Choi, S.R. 2014 Properties of desert sand and CMAS glass. Tech. Rep. NASA/TM-2014-218365. NASA Glenn Research Center Cleveland.Google Scholar
Billard, L. & Diday, E. 2002 Symbolic regression analysis. In Classification, Clustering, and Data Analysis: Recent Advances and Applications (ed. K. Jajuga, A. Sokołowski & H.-H. Bock), pp. 281–288. Springer.CrossRefGoogle Scholar
Bishop, C.M. 1997 Bayesian neural networks. J. Braz. Comput. Soc. 4, 6168.CrossRefGoogle Scholar
Blei, D.M., Kucukelbir, A. & McAuliffe, J.D. 2017 Variational inference: a review for statisticians. J. Am. Stat. Assoc. 112 (518), 859877.CrossRefGoogle Scholar
Bonn, D., Eggers, J., Indekeu, J., Meunier, J. & Rolley, E. 2009 Wetting and spreading. Rev. Mod. Phys. 81 (2), 739805.CrossRefGoogle Scholar
Brooks, S. 1998 Markov Chain Monte Carlo method and its application. J. R. Stat. Soc. D 47, 69100.Google Scholar
Brunton, S.L., Proctor, J.L. & Kutz, J.N. 2016 Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proc. Natl Acad. Sci. USA 113 (15), 39323937.CrossRefGoogle ScholarPubMed
Bykov, K., Höhne, M.M.-C., Creosteanu, A., Müller, K.-R., Klauschen, F., Nakajima, S. & Kloft, M. 2021 Explaining Bayesian neural networks. arXiv:2108.10346.Google Scholar
Chan, K.C., Li, Z. & Wenzel, W. 2023 A Mori–Zwanzig dissipative particle dynamics approach for anisotropic coarse grained molecular dynamics. J. Chem. Theory Comput. 19 (3), 910923.CrossRefGoogle ScholarPubMed
Chen, L., Bonaccurso, E., Deng, P. & Zhang, H. 2016 Droplet impact on soft viscoelastic surfaces. Phys. Rev. E 94, 063117.CrossRefGoogle ScholarPubMed
Chen, Y., Lu, L., Karniadakis, G.E. & Dal Negro, L. 2020 Physics-informed neural networks for inverse problems in nano-optics and metamaterials. Opt. Express 28, 1161811633.CrossRefGoogle ScholarPubMed
Clarke, D.R., Oechsner, M. & Padture, N.P. 2012 Thermal-barrier coatings for more efficient gas-turbine engines. MRS Bull. 37 (10), 891898.CrossRefGoogle Scholar
Cormier, S.L., McGraw, J.D., Salez, T., Raphaël, E. & Dalnoki-Veress, K. 2012 Beyond Tanner's law: crossover between spreading regimes of a viscous droplet on an identical film. Phys. Rev. Lett. 109 (15), 154501.CrossRefGoogle Scholar
De Ruijter, M.J., De Coninck, J. & Oshanin, G. 1999 Droplet spreading: partial wetting regime revisited. Langmuir 15 (6), 22092216.CrossRefGoogle Scholar
Delahunt, C.B. & Kutz, J.N. 2022 A toolkit for data-driven discovery of governing equations in high-noise regimes. IEEE Access 10, 3121031234.CrossRefGoogle Scholar
Dussan, E.B. 1979 On the spreading of liquids on solid surfaces: static and dynamic contact lines. Annu. Rev. Fluid Mech. 11 (1), 371400.CrossRefGoogle Scholar
Eddi, A., Winkels, K.G. & Snoeijer, J.H. 2013 Short time dynamics of viscous drop spreading. Phys. Fluids 25 (1), 013102.CrossRefGoogle Scholar
Edwards, A.M.J., Ledesma-Aguilar, R., Newton, M.I., Brown, C.V. & McHale, G. 2020 A viscous switch for liquid–liquid dewetting. Commun. Phys. 3 (1), 16.CrossRefGoogle Scholar
Español, P. & Warren, P. 1995 Statistical mechanics of dissipative particle dynamics. Europhys. Lett. 30 (4), 191.CrossRefGoogle Scholar
Español, P. & Warren, P.B. 2017 Perspective: dissipative particle dynamics. J. Chem. Phys. 146 (15), 150901.CrossRefGoogle ScholarPubMed
de Gennes, P.G. 1985 Wetting: statics and dynamics. Rev. Mod. Phys. 57 (3), 827863.CrossRefGoogle Scholar
Ghoufi, A. & Malfreyt, P. 2012 Coarse grained simulations of the electrolytes at the water–air interface from many body dissipative particle dynamics. J. Chem. Theory Comput. 8 (3), 787791.CrossRefGoogle ScholarPubMed
Gordillo, J.M., Riboux, G. & Quintero, E.S. 2019 A theory on the spreading of impacting droplets. J. Fluid Mech. 866, 298315.CrossRefGoogle Scholar
Gorin, B., Di Mauro, G., Bonn, D. & Kellay, H. 2022 Universal aspects of droplet spreading dynamics in Newtonian and non-Newtonian fluids. Langmuir 38 (8), 26082613.CrossRefGoogle ScholarPubMed
Grant, K.M., Krämer, S., Löfvander, J.P.A. & Levi, C.G. 2007 CMAS degradation of environmental barrier coatings. Surf. Coat. Technol. 202 (4–7), 653657.CrossRefGoogle Scholar
Graves, A. 2011 Practical variational inference for neural networks. In Advances in Neural Information Processing Systems (ed. J. Shawe-Taylor, R. Zemel, P. Bartlett, F. Pereira & K.Q. Weinberger), vol. 24, pp. 2348–2356. Curran Associates.Google Scholar
Groot, R.D. 2004 Applications of dissipative particle dynamics. In Novel Methods in Soft Matter Simulations (ed. M. Karttunen, A. Lukkarinen & I. Vattulainen), pp. 5–38. Springer.CrossRefGoogle Scholar
Hardy, W.B. 1919 III. The spreading of fluids on glass. Lond. Edinb. Dublin Philos. Mag. J. Sci. 38 (223), 4955.CrossRefGoogle Scholar
Hassan, G., Yilbas, B.S., Al-Sharafi, A. & Al-Qahtani, H. 2019 Self-cleaning of a hydrophobic surface by a rolling water droplet. Sci. Rep. 9, 114.CrossRefGoogle ScholarPubMed
Jain, N., Lemoine, A., Chaussonnet, G., Flatau, A., Bravo, L., Ghoshal, A., Walock, M. & Murugan, M. 2021 A critical review of physical models in high temperature multiphase fluid dynamics: turbulent transport and particle-wall interactions. Appl. Mech. Rev. 73, 040801.CrossRefGoogle Scholar
Karniadakis, G.E., Kevrekidis, I.G., Lu, L., Perdikaris, P., Wang, S. & Yang, L. 2021 Physics-informed machine learning. Nat. Rev. Phys. 3, 422440.CrossRefGoogle Scholar
Kiyani, E., Shukla, K., Karniadakis, G.E. & Karttunen, M. 2023 A framework based on symbolic regression coupled with extended physics-informed neural networks for gray-box learning of equations of motion from data. arXiv:2305.10706.CrossRefGoogle Scholar
Kiyani, E., Silber, S., Kooshkbaghi, M. & Karttunen, M. 2022 Machine-learning-based data-driven discovery of nonlinear phase-field dynamics. Phys. Rev. E 106, 065303.CrossRefGoogle ScholarPubMed
Koneru, R.B., Flatau, A., Li, Z., Bravo, L., Murugan, M., Ghoshal, A. & Karniadakis, G.E. 2022 Quantifying the dynamic spreading of a molten sand droplet using multiphase mesoscopic simulations. Phys. Rev. Fluids 7 (10), 103602.CrossRefGoogle Scholar
de Laplace, P.-S. 1805 Supplément au livre X du traité de mécanique céleste. Sur l'action capillaire. In Traité de mécanique céleste. Gauthier-Vilars.Google Scholar
Lee, S., Kooshkbaghi, M., Spiliotis, K., Siettos, C.I. & Kevrekidis, I.G. 2020 Coarse-scale PDEs from fine-scale observations via machine learning. Chaos 30, 013141.CrossRefGoogle ScholarPubMed
Lei, L., Bertevas, E.L., Khoo, B.C. & Phan-Thien, N. 2018 Many-body dissipative particle dynamics (MDPD) simulation of a pseudoplastic yield-stress fluid with surface tension in some flow processes. J. Non-Newtonian Fluid Mech. 260, 163174.CrossRefGoogle Scholar
Li, Z., Bian, X., Tang, Y.-H. & Karniadakis, G.E. 2018 A dissipative particle dynamics method for arbitrarily complex geometries. J. Comput. Phys. 355, 534547.CrossRefGoogle Scholar
Li, Z., Bian, X., Yang, X. & Karniadakis, G.E. 2016 A comparative study of coarse-graining methods for polymeric fluids: Mori–Zwanzig vs iterative Boltzmann inversion vs stochastic parametric optimization. J. Chem. Phys. 145 (4), 044102.CrossRefGoogle ScholarPubMed
Li, Z., Hu, G.-H., Wang, Z.-L., Ma, Y.-B. & Zhou, Z.-W. 2013 Three dimensional flow structures in a moving droplet on substrate: a dissipative particle dynamics study. Phys. Fluids 25, 072103.CrossRefGoogle Scholar
Lucy, L.B. 1977 A numerical approach to the testing of the fission hypothesis. Astron. J. 82, 1013.CrossRefGoogle Scholar
Mao, Z., Jagtap, A.D. & Karniadakis, G.E. 2020 Physics-informed neural networks for high-speed flows. Comput. Meth. Appl. Mech. Engng 360, 112789.CrossRefGoogle Scholar
Martius, G. & Lampert, C.H. 2016 Extrapolation and learning equations. arXiv:1610.02995.Google Scholar
McGraw, J.D., Chan, T.S., Maurer, S., Salez, T., Benzaquen, M., Raphaël, E., Brinkmann, M. & Jacobs, K. 2016 Slip-mediated dewetting of polymer microdroplets. Proc. Natl Acad. Sci. USA 113 (5), 11681173.CrossRefGoogle ScholarPubMed
McHale, G., Shirtcliffe, N.J., Aqil, S., Perry, C.C. & Newton, M.I. 2004 Topography driven spreading. Phys. Rev. Lett. 93 (3), 036102.CrossRefGoogle ScholarPubMed
Meidani, K. & Farimani, A.B. 2021 Data-driven identification of 2D partial differential equations using extracted physical features. Comput. Meth. Appl. Mech. Engng 381, 113831.CrossRefGoogle Scholar
Mishra, S. & Molinaro, R. 2020 Estimates on the generalization error of physics informed neural networks (PINNs) for approximating PDEs. II. A class of inverse problems. arXiv:2007.01138.CrossRefGoogle Scholar
Murtola, T., Bunker, A., Vattulainen, I., Deserno, M. & Karttunen, M. 2009 Multiscale modeling of emergent materials: biological and soft matter. Phys. Chem. Chem. Phys. 11 (12), 18691892.CrossRefGoogle ScholarPubMed
Naraparaju, R., Gomez Chavez, J.J., Niemeyer, P., Hess, K.-U., Song, W., Dingwell, D.B., Lokachari, S., Ramana, C.V. & Schulz, U. 2019 Estimation of CMAS infiltration depth in EB-PVD TBCs: a new constraint model supported with experimental approach. J. Eur. Ceram. Soc. 39 (9), 29362945.CrossRefGoogle Scholar
Ndamka, N.L., Wellman, R.G. & Nicholls, J.R. 2016 The degradation of thermal barrier coatings by molten deposits: introducing the concept of basicity. Mater. High Temp. 33 (1), 4450.CrossRefGoogle Scholar
Neal, R.M. 2011 MCMC using Hamiltonian dynamics. In Handbook of Markov Chain Monte Carlo (ed. S. Brooks, A. Gelman, G. Jones & X.-L. Meng), pp. 113–162. Chapman & Hall/CRC.CrossRefGoogle Scholar
Neal, R.M. 2012 Bayesian Learning for Neural Networks, vol. 118. Springer Science & Business Media.Google Scholar
Nieminen, J.A., Abraham, D.B., Karttunen, M. & Kaski, K. 1992 Molecular dynamics of a microscopic droplet on solid surface. Phys. Rev. Lett. 69 (1), 124127.CrossRefGoogle ScholarPubMed
Nieto, A., Agrawal, R., Bravo, L., Hofmeister-Mock, C., Pepi, M. & Ghoshal, A. 2021 Calcia–Magnesia–Alumina–Silicate (CMAS) attack mechanisms and roadmap towards sandphobic thermal and environmental barrier coatings. Intl Mater. Rev. 66 (7), 451492.CrossRefGoogle Scholar
Nishimoto, S. & Bhushan, B. 2013 Bioinspired self-cleaning surfaces with superhydrophobicity, superoleophobicity, and superhydrophilicity. RSC Adv. 3 (3), 671690.CrossRefGoogle Scholar
Pitois, O. & François, B. 1999 Crystallization of condensation droplets on a liquid surface. Colloid Polym. Sci. 277, 574578.CrossRefGoogle Scholar
Poerschke, D.L. & Levi, C.G. 2015 Effects of cation substitution and temperature on the interaction between thermal barrier oxides and molten CMAS. J. Eur. Ceram. Soc. 35 (2), 681691.CrossRefGoogle Scholar
Popescu, M.N., Oshanin, G., Dietrich, S. & Cazabat, A.-M. 2012 Precursor films in wetting phenomena. J. Phys.: Condens. Matter 24 (24), 243102.Google ScholarPubMed
Radivojević, T. & Akhmatskaya, E. 2020 Modified Hamiltonian Monte Carlo for Bayesian inference. Stat. Comput. 30 (2), 377404.CrossRefGoogle Scholar
Raissi, M., Perdikaris, P. & Karniadakis, G.E. 2019 Physics-informed neural networks: a deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations. J. Comput. Phys. 378, 686707.CrossRefGoogle Scholar
Rao, Q., Xia, Y., Li, J., McConnell, J., Sutherland, J. & Li, Z. 2021 A modified many-body dissipative particle dynamics model for mesoscopic fluid simulation: methodology, calibration, and application for hydrocarbon and water. Mol. Simul. 47 (4), 363375.CrossRefGoogle Scholar
Ren, J. & Duan, J. 2020 Identifying stochastic governing equations from data of the most probable transition trajectories. arXiv:2002.10251.Google Scholar
Shukla, K., Di Leoni, P.C., Blackshire, J., Sparkman, D. & Karniadakis, G.E. 2020 Physics-informed neural network for ultrasound nondestructive quantification of surface breaking cracks. J. Nondestruct. Eval. 39, 120.CrossRefGoogle Scholar
Song, W., Lavallée, Y., Hess, K.-U., Kueppers, U., Cimarelli, C. & Dingwell, D.B. 2016 Volcanic ash melting under conditions relevant to ash turbine interactions. Nat. Commun. 7, 10795.CrossRefGoogle Scholar
Srinivasan, V. & Mason, C.H. 1986 Nonlinear least squares estimation of new product diffusion models. Market. Sci. 5 (2), 169178.CrossRefGoogle Scholar
Stephens, T. 2016 Genetic programming in Python, with a scikit-learn inspired API: gplearn.Google Scholar
Tanner, L.H. 1979 The spreading of silicone oil drops on horizontal surfaces. J. Phys. D: Appl. Phys. 12 (9), 1473.CrossRefGoogle Scholar
Thiem, T.N., Kooshkbaghi, M., Bertalan, T., Laing, C.R. & Kevrekidis, I.G. 2020 Emergent spaces for coupled oscillators. Front. Comput. Neurosci. 14, 36.CrossRefGoogle ScholarPubMed
Thompson, A.P., et al. 2022 LAMMPS – a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Comms. 271, 108171.CrossRefGoogle Scholar
Vidal-Setif, M.H., Chellah, N., Rio, C., Sanchez, C. & Lavigne, O. 2012 Calcium–magnesium– alumino-silicate (CMAS) degradation of EB-PVD thermal barrier coatings: characterization of CMAS damage on ex-service high pressure blade TBCs. Surf. Coat. Technol. 208, 3945.CrossRefGoogle Scholar
Warren, P.B. 2001 Hydrodynamic bubble coarsening in off-critical vapor–liquid phase separation. Phys. Rev. Lett. 87 (22), 225702.CrossRefGoogle ScholarPubMed
Warren, P.B. 2003 Vapor–liquid coexistence in many-body dissipative particle dynamics. Phys. Rev. E 68 (6 Pt 2), 066702.CrossRefGoogle ScholarPubMed
Wiesner, V.L., Vempati, U.K. & Bansal, N.P. 2016 High temperature viscosity of calcium-magnesium- aluminosilicate glass from synthetic sand. Scr. Mater. 124, 189192.CrossRefGoogle Scholar
Winkels, K.G., Weijs, J.H., Eddi, A. & Snoeijer, J.H. 2012 Initial spreading of low-viscosity drops on partially wetting surfaces. Phys. Rev. E 85, 055301.CrossRefGoogle ScholarPubMed
Xia, Y., Goral, J., Huang, H., Miskovic, I., Meakin, P. & Deo, M. 2017 Many-body dissipative particle dynamics modeling of fluid flow in fine-grained nanoporous shales. Phys. Fluids 29 (5), 056601.CrossRefGoogle Scholar
Yang, L., Meng, X. & Karniadakis, G.E. 2021 B-PINNs: Bayesian physics-informed neural networks for forward and inverse PDE problems with noisy data. J. Comput. Phys. 425, 109913.CrossRefGoogle Scholar
Young, T. 1805 An essay on the cohesion of fluids. Phil. Trans. R. Soc. Lond. 95, 6587.Google Scholar
Zhao, J., Chen, S., Zhang, K. & Liu, Y. 2021 A review of many-body dissipative particle dynamics (MDPD): theoretical models and its applications. Phys. Fluids 33 (11), 112002.CrossRefGoogle Scholar
Figure 0

Figure 1. A schematic showing the equilibrium contact angle (i.e. $\theta \equiv \theta _{eq}$), the surface tensions ($\gamma$), the threshold between low- and high-wetting regimes ($\theta _{eq} =90^{\circ }$), and a situation of a non-wetting droplet ($\theta _{eq} =180^{\circ }$). The last image demonstrates the occurrence of a precursor that is observed in some cases. In that case, the (macroscopic) contact angle is defined using the macroscopic part of the droplet, the cap in the rightmost image. The height of the precursor is in the molecular length scales (Hardy 1919; Nieminen et al.1992; Popescu et al.2012).

Figure 1

Figure 2. The equilibrium contact angles $\theta _{eq}$ for the different attraction parameters between the liquid and solid particles ($A_{ls}$; see (2.7)). It is worth noting that the data for this figure have been extracted from Koneru et al. (2022).

Figure 2

Figure 3. (a) Illustration of the spreading behaviour of a CMAS droplet on a hydrophilic surface at different times. The droplet with initial size $R_0$ spreads on the surface with radius $r(t)$ and contact angle $\theta (t)$. (b) A series of snapshots from a simulation of a droplet with initial size $R_0 = 0.136$ mm and equilibrium contact angle $\theta _{eq} = 93.4^\circ$.

Figure 3

Figure 4. The impact of $\theta _{eq}$ and $R_0$ on the droplet radii as a function of time for various (a) initial drop sizes with equilibrium contact angle $\theta _{eq} = 54.6^\circ$ corresponding to $A_{ls} = 30.0$, and (b) equilibrium contact angles (corresponding to $A_{ls} = -25.0, -25.8, -27.0, -28.0, -29.0, -30.0, -31.4, -32.2$) and initial drop size $R_{0}=0.136$ mm.

Figure 4

Figure 5. The value of $\alpha$, calculated using (1.2), varies for different initial radii and fixed equilibrium contact angles (a) $\theta _{eq}=85.6^\circ$ and (b) $\theta _{eq}=62.4^\circ$. The figure illustrates that $\alpha$ is influenced by both the initial drop size $R_0$ and the equilibrium contact angle $\theta _{eq}$.

Figure 5

Figure 6. The process of utilizing PINNs to extract three unknown parameters of the ODE (3.1), using three-dimensional mDPD simulation data. First, a neural network is trained using simulation data, where the input is time $t$ and the output is spreading radii $\tilde {r}(t)$. This neural network comprises four layers with three neurons, and is trained for 12 000 epochs. Subsequently, the predicted $\tilde {r}(t)$ is used to satisfy (3.1) in the physics-informed part. The loss function for this process consists of two parts: data matching and residual. By optimizing the loss function, the values of $\eta (R_0, \theta _{eq})$, $\beta (R_0, \theta _{eq})$ and $\tau (R_0,\theta _{eq})$ are determined for each set of $R_0$ and $\theta _{eq}$. After predicting the unknown parameters using PINNs, two additional neural networks, denoted as $NN_\beta$ and $NN_\tau$, are trained using these parameters to generate values for the unknown parameters at points where data are not available. The outputs of these networks, together with the outputs of the PINNs, are then fed through a symbolic regression model to discover a mathematical expression for the discovered parameter.

Figure 6

Figure 7. Comparison of the time evolution of the droplet radii: mDPD simulations (symbols), ODE model (3.1) (solid lines) and PINN predictions (dashed lines) for $\theta _{eq}=\{39.1^\circ, 62.4^\circ,93.4^\circ \}$ and $R_0=\{0.136, 0.153, 0.187,0.204\}$ mm parameter sets.

Figure 7

Figure 8. (a,b,c) The evolution of parameters $\eta (R_0,\theta _{eq})$, $\beta (R_0,\theta _{eq})$ and $\tau (R_0,\theta _{eq})$, respectively, over multiple epochs. These plots demonstrate that the parameters gradually converge to a stable state after $12\,000$ epochs. (d) The traces of the loss function for the PINNs framework. The learning curves demonstrate the decreasing trend of the loss functions, indicating that they converge to a stable point for all initial drop sizes and $\theta _{eq}$.

Figure 8

Figure 9. The values of (a) $\eta$, (b) $\beta$ and (c) $\tau$ obtained through PINNs. These values exhibit varying behaviours depending on the initial radius $R_{0}$ and equilibrium contact angle $\theta _{eq}$. The horizontal axes display the equilibrium contact angles $\theta _{eq}$. The vertical axes of all plots represent the values of $\eta$, $\beta$ and $\tau$. Here, $\eta$ remains nearly constant within a small range of values between $-0.325$ and $-0.200$, and $\beta$ and $\tau$ change within ranges from $1.0$ to $5.0$, and $6.5$ to $8.0$, respectively.

Figure 9

Figure 10. Predictions of the parameters (a) $\tau$ and (b) $\beta$ using the trained neural networks $NN_{\beta }$ and $NN_{\tau }$. The horizontal axes show $R_0$ and $\theta _{eq}$. The green and red circles correspond to the obtained values of $\tau$ and $\beta$ using the PINNs that were used to train $NN_{\beta }$ and $NN_{\tau }$. Additionally, the orange and blue dots represent the predicted values for grid interpolations from $R_0=0.136\ {\rm mm}$ to $R_0=0.204\ {\rm mm}$, and $\theta _{eq}=40^\circ$ to $\theta _{eq}=95^\circ$.

Figure 10

Figure 11. The behaviour of $\alpha$ from the right-hand side of (3.1) with parameters from (5.1): (a) different contact angles with fixed initial radius $R_0=0.136$ mm; (b) varying initial radii with fixed contact angle $\theta _{eq}=77.9^\circ$.

Figure 11

Figure 12. (a) The behaviour of the parameter $\alpha$ using (3.1) for $R_{0} = 0.127$ mm, which falls outside the range of the initial drop sizes used for training the networks. (b) The simulation data and the solution obtained from solving the ODE (3.1) with parameters from symbolic regression (5.1).

Figure 12

Figure 13. The mean and uncertainty (mean $\pm$2 standard deviations) of B-PINN predictions of the spreading radii history are given as solid lines and shaded regions, respectively. The test simulation data are depicted by solid circles, and training data are indicated by stars. This analysis is carried out for two different initial drop sizes, namely $R_0 = 0.136$ and $0.170$ mm, for three equilibrium contact angles.

Figure 13

Figure 14. Comparison between B-PINNs and PINNs discovered parameters for a range of equilibrium contact angles and two initial radii. The mean values (solid lines) and the standard deviations (mean values $\pm$2 standard deviations, shaded region) of (a,c) $\beta$ and (b,d) $\tau$. The dashed lines represent the parameters discovered by PINNs.

Figure 14

Figure 15. Comparison between the ODE solution with parameters found by B-PINNs (solid lines), and the simulation radii (circles). Two initial drop sizes, $R_0 = 0.136$ and $0.170$ mm, and three equilibrium contact angles are shown.