1. Introduction
Understanding droplet impact on solid surfaces is critical for various engineering applications, including spray cooling, coating technologies and ice accretion in aeronautical systems. In particular, in icing conditions, droplet impact is the precursor to ice formation, making accurate heat transfer modelling essential for predicting ice growth. This study is motivated by the need to improve thermal modelling in such applications, where droplet impact dictates the initial thermal exchange at the interface. Given its significance in both fundamental research and practical applications, droplet impact has been extensively studied over the years. The interest in this phenomenon dates back nearly one and a half centuries to the pioneering observations of Worthington (Reference Worthington1877). As noted by Thoroddsen, Etoh & Takehara (Reference Thoroddsen, Etoh and Takehara2008), the advent of high-speed imaging techniques has greatly increased interest in this field, providing deeper insights over recent decades. This interest spans multiple fields and applications; for instance, the control and understanding of droplet impact dynamics have led to advancements in spray-based technologies, such as fuel injection in internal combustion engines, as reviewed by Moreira, Moita & Pan (Reference Moreira, Moita and Panão2010). In forensic science, blood-droplet impact patterns have been analysed to reconstruct events, as discussed by Adam (Reference Adam2012), Brodbeck (Reference Brodbeck2012) and Laan et al. (Reference Laan, de Bruin, Bartolo, Josserand and Bonn2014). Surface engineering has also benefited from research on droplet interactions with complex substrates, as examined by Marengo et al. (Reference Marengo, Antonini, Roisman and Tropea2011). In icing applications, the thermal exchange initiated by droplet impacts governs the onset of ice buildup on aircraft and power-infrastructure surfaces, as reviewed in Roisman & Tropea (Reference Roisman and Tropea2021). In the aerospace and renewable energy sectors, the ability to control droplet impact is crucial for minimising substrate damage. For instance, Keegan, Nash & Stack (Reference Keegan, Nash and Stack2013) studied the erosion of wind turbines, while Zhang et al. (Reference Zhang, Ma, Liu, Ren and Hu2021) analysed the degradation of protective coatings. These diverse applications highlight the broad relevance of understanding droplet impact across multiple disciplines.
Dimensionless parameters governing the drop impact dynamics and the phenomena associated with drop spreading and splashing include the Reynolds
${\textit{Re}}$
and Weber
$We$
numbers, which are defined as
where
$D$
and
$U_{{d,i}}$
are the drop initial diameter and impact velocity,
$ \mu _{{d}}$
,
$ \rho _{{d}}$
and
$ \sigma _{{d}}$
denote the dynamic viscosity, density and surface tension of the liquid drop, respectively. The subscript
$ i$
distinguishes whether the parameter is based on the normal, n, or tangential, t, relative velocity component between the droplet and the substrate. The Reynolds number expresses the ratio of inertial to viscous forces in the spreading film, while the Weber number represents the ratio of inertial to capillary forces.
Over the decades, Rein (Reference Rein2002), Yarin (Reference Yarin2006), Josserand & Thoroddsen (Reference Josserand and Thoroddsen2016) and Mohammad Karim (Reference Mohammad Karim2023) have provided comprehensive reviews summarising progress in understanding droplet impact. Early studies primarily focused on cases involving high Reynolds and Weber numbers with single droplets.
For example, Yarin & Weiss (Reference Yarin and Weiss1995) investigated periodic droplet trains impacting substrates and extended theoretical insights to single-droplet impacts on dry surfaces. Their approach introduced a self-similar solution for the internal flow during spreading, which laid the groundwork for subsequent investigations. Building on this foundation, Roisman (Reference Roisman2009) and Roisman et al. (Reference Roisman, Berberović and Tropea2009) refined self-similar solutions of the Navier–Stokes equations to predict the maximum spreading diameter and the residual film thickness under isothermal conditions. Subsequently, Riboux & Gordillo (Reference Riboux and Gordillo2014) developed theories on the critical impact speed required for drop splashing, which they expanded in Riboux & Gordillo (Reference Riboux and Gordillo2015) to quantify the mass and velocities of the ejected droplets. For asymmetric spreading, García-Geijo et al. (Reference García-Geijo, Riboux and Gordillo2020), drawing on the theoretical framework of Gordillo, Riboux & Quintero (Reference Gordillo, Riboux and Quintero2019), predicted the time-evolving shape of the border of the thin liquid film. Xu, Zhang & Nagel (Reference Xu, Zhang and Nagel2005) investigated how the surrounding gas pressure affects the stresses and flow fields within the spreading droplet. Additionally, Stow & Hadfield (Reference Stow and Hadfield1981), Range & Feuillebois (Reference Range and Feuillebois1998), Xu (Reference Xu2007), Latka et al. (Reference Latka, Strandburg-Peshkin, Driscoll, Stevens and Nagel2012), Quetzeri-Santiago, Castrejón-Pita & Castrejón-Pita (Reference Quetzeri-Santiago, Castrejón-Pita and Castrejón-Pita2019) and García-Geijo et al. (Reference García-Geijo, Quintero, Riboux and Gordillo2021) studied the effects of substrate roughness, while Bartolo, Josserand & Bonn (Reference Bartolo, Josserand and Bonn2005), Quéré (Reference Quéré2008) and Antonini, Amirfazli & Marengo (Reference Antonini, Amirfazli and Marengo2012) emphasised the critical role of wettability. Studies by Antonini, Villa & Marengo (Reference Antonini, Villa and Marengo2014), Yeong et al. (Reference Yeong, Burton, Loth and Bayer2014), LeClear et al. (Reference LeClear, LeClear, Abhijeet, Park and Choi2016), Regulagadda, Bakshi & Das (Reference Regulagadda, Bakshi and Das2018) and Aboud & Kietzig (Reference Aboud and Kietzig2018) also considered non-axisymmetric impacts and examined the performance of specialised substrates, such as superhydrophobic and icephobic materials, to enhance anti-icing solutions.
Substrate motion introduces an additional complexity, as investigated by Mundo, Sommerfeld & Tropea (Reference Mundo, Sommerfeld and Tropea1995), Almohammadi & Amirfazli (Reference Almohammadi and Amirfazli2017) and Buksh et al. (Reference Buksh, Almohammadi, Marengo and Amirfazli2019). Inclined-impact configurations can replicate certain aspects of a moving substrate at low velocities, as demonstrated by Buksh, Marengo & Amirfazli (Reference Buksh, Marengo and Amirfazli2020). However, Povarov et al. (Reference Povarov, Nazarov, Ignat’evskaya and Nikol’skii1976) and Stumpf et al. (Reference Stumpf, Qezeljeh, Kamal, Dezitter, Martuffo, Roisman and Hussong2024) reported that sufficiently high substrate speeds generate boundary layers strong enough to induce aerodynamic rebound, resulting in dynamics that do not occur in inclined-impact scenarios.
Although these investigations provided profound insights into isothermal conditions, incorporating thermal effects such as heating, cooling or phase change is still challenging. For example, when the substrate is heated, Senda et al. (Reference Senda, Yamada, Fujimoto and Miki1988) and Bernardin et al. (Reference Bernardin, Stebbins and Mudawar1997) demonstrated that phenomena like the Leidenfrost effect can significantly alter flow patterns and breakup modes. Understanding these transitions is critical for applications such as spray cooling of hot surfaces, as emphasised by Breitenbach, Roisman & Tropea (Reference Breitenbach, Roisman and Tropea2018). Conversely, when a droplet solidifies upon impact, its morphology evolves differently. Such scenarios arise in processes including spray forming, as Orme (Reference Orme1993) described, plasma spray coating, studied by Fauchais (Reference Fauchais2004), thin-film fabrication, as noted by Steirer et al. (Reference Steirer, Reese, Rupert, Kopidakis, Olson, Collins and Ginley2009), and ice accretion in aeronautical applications, as discussed by Cao, Tan & Wu (Reference Cao, Tan and Wu2018). Accurate characterisation of thermal behaviour at the droplet–substrate interface is essential for predicting and analysing droplet impact outcomes. Tarozzi, Muscio & Tartarini (Reference Tarozzi, Muscio and Tartarini2007), Teodori et al. (Reference Teodori, Pontes, Moita and Moreira2018), Schmidt et al. (Reference Schmidt, Breitenbach, Roisman and Tropea2020), Gholijani et al. (Reference Gholijani, Schlawitschek, Gambaryan-Roisman and Stephan2020) and Gholijani et al. (Reference Gholijani, Fischer, Gambaryan-Roisman and Stephan2022) demonstrated that infrared thermography is a highly effective technique for capturing rapidly changing temperature fields during impact. In this study, we utilise an infrared-transparent substrate with an appropriate coating to obtain detailed, two-dimensional (2-D) temperature fields at high temporal resolution, enabling precise measurements of contact temperature distributions and heat fluxes.
Thermal modelling of drop impact has traditionally been based on simplified frameworks since the real problem is rather complicated. For instance, Seki, Kawamura & Sanokawa (Reference Seki, Kawamura and Sanokawa1978) proposed for the description of the heat transfer in a droplet suddenly coming into contact with a solid wall to use a one-dimensional (1-D) model for heat conduction in two semi-infinite solid regions (Carslaw & Jaeger Reference Carslaw and Jaeger1959). The expressions for the contact temperature
$T_c$
and the heat flux at the interface
$q_c$
are obtained in the form
Here,
$T_{{d}}$
and
$T_{{w}}$
represent the temperatures of the liquid and solid, respectively, while
$e_{{d}}$
and
$e_{{w}}$
denote their thermal effusivities.
To improve modelling, some studies introduce a finite interfacial resistance
$R_c$
to represent the temperature jump and to regularise the early-time
$t^{-1/2}$
heat-flux growth (Aziz & Chandra Reference Aziz and Chandra2000; Chandra & Fauchais Reference Chandra and Fauchais2009). In parallel, to account for the flow inside the drop, Roisman (Reference Roisman2009) derived a self-similar spreading-flow solution that Roisman (Reference Roisman2010) used to include convective heat transport, although interfacial effects associated with substrate morphology were not considered,
The term
$\mathscr{I}$
on the right-hand side of (1.3) accounts for the influence of convective effects in the liquid phase and is a function of the liquid Prandtl number
$ {\textit{Pr}}_{{d}}$
, computed in Roisman (Reference Roisman2010).
Although several studies have measured contact temperatures on stationary substrates, no corresponding data for moving substrates are available in the existing literature. The validity of the existing models requires an accurate experimental evaluation and validation.
This study has been primarily motivated by the need for improved heat transfer modelling in applications where liquid droplet impact plays a crucial role in ice accretion, setting the stage for subsequent freezing. Specifically, we examine how droplet diameter, substrate speed and thermal boundary conditions influence contact temperatures and heat fluxes. Our theoretical approach builds on the solution obtained in Carslaw & Jaeger (Reference Carslaw and Jaeger1959), incorporating the convective effect related to the velocity of the spreading drop. These developments are complemented by high-resolution experiments using high-speed imaging and infrared thermography, providing a comprehensive understanding of the flow and thermal fields during the early stages of droplet impact on a moving substrate.
2. Experimental study
2.1. Experimental set-up
An experimental set-up was developed to investigate the heat transfer mechanisms during droplet impact. The configuration is schematically illustrated in figure 1. Droplets of double-distilled water are generated within a temperature-controlled cooling chamber (marked 4 in figure 1), ensuring consistent and well-defined initial and ambient conditions. The temperature of the droplets is measured using a type-T thermocouple, marked 5, with an accuracy of
$\pm 0.5\,^\circ$
C, positioned inside the droplet-generation needle. The cooling system, marked 3, ensures precise control of the droplet’s initial temperature. The moving substrate consists of a graphite-coated sapphire disc, marked 8, that serves as the target surface. This substrate is rotated using a geared motor, marked 9, with precise speed control, allowing for accurate setting of the angular velocity. A magnetic encoder, marked 10, is used to provide feedback on the disc’s rotation, which is managed by a motor driver controller, marked 11. To synchronise the measurements, a data acquisition system, marked 16, is used, while environmental parameters, such as humidity and temperature, are monitored using a dedicated sensor, marked 13, with data retrieved by a microcontroller, marked 14. The Sensirion SHT85 sensor measures temperature with
$\pm 0.1\,^\circ$
C accuracy and relative humidity with
$\pm$
1.5 % error. The personal computer, marked 15, is used for data logging and processing. The impact radius,
$ R$
, is fixed at
$ 73 \pm 1$
mm, such that no edge effects are present in the boundary layer (Pier Reference Pier2013). A high-speed camera, marked 2, positioned laterally to the substrate, records the impact dynamics, enabling measurement of the droplet’s normal velocity relative to the substrate,
$ U_{{d,n}}$
, and its diameter,
$ D$
. The Phantom V2012 high-speed camera operates at 98 900 Hz, with spatial resolution of either 30.9 or 26.7
$\unicode{x03BC} \textrm{m}\,\textrm{pixel}^{-1}$
, depending on the experiment. A high-speed infrared camera, marked 1, is positioned orthogonally to the substrate’s axis of rotation to capture the thermal interactions during droplet impact. The Telops FAST M3k infrared camera covers a spectral range of 1.5–5.4
$\unicode{x03BC}$
m, operates at 2300–8600 Hz, and has a spatial resolution of 73 or 200
$\unicode{x03BC} \textrm{m}\,\textrm{pixel}^{-1}$
, depending on the experiment.

Figure 1. Illustration of the experimental set-up developed to investigate heat transfer during droplet impact. The set-up comprises a droplet-generation system, a rotating sapphire disc as the substrate, high-speed cameras for visualising droplet dynamics, an infrared camera for capturing temperature fields and environmental monitoring equipment.

Figure 2. Representative images of droplet impact on a moving substrate. Panels (a) and (b) show the droplet impact at a substrate velocity of 0.38 m s−1, with a droplet diameter
$ D = 2.45 \, \textrm{mm}$
, normal impact velocity
$ U_{{d,n}} = 3.13 \, {\textrm{m s}^{- 1}}$
, initial droplet temperature
$ T_{{d,0}} = 0.5\,^\circ \textrm{C}$
and initial substrate temperature
$ T_{{w,0}} = 26.1\,^\circ \textrm{C}$
. Panels (c) and (d) depict the droplet impact at a substrate velocity of 6.12 m s−1 under otherwise identical conditions. An asymmetry in the droplet spreading is visible, arising from the increased substrate motion.
Exemplary side-view images of the spreading drop are shown in figure 2. These were captured using the high-speed video system, marked 2. Thermocouple measurements are used to extract the corresponding temperature-dependent properties of water from VDI (Reference VDI2013). In the laboratory reference frame, the droplet possesses no tangential velocity. Consequently, its tangential velocity relative to the substrate equals the substrate speed,
$ \omega R = U_{{d,t}}$
, where
$ \omega$
is the angular velocity of the substrate, derived from the motor’s rotational speed,
$ n$
.
The Reynolds number for the airflow induced by the moving substrate is denoted by
${\textit{Re}}_{{s}} = \omega R^2/\nu _{{a}}$
, where
$ \nu _{{a}}$
is the kinematic viscosity of air. In these experiments,
$ {Re}_{{s}}$
ranges from 2000 to 30 000. According to Kobayashi (Reference Kobayashi1994), the critical Reynolds number for the onset of instabilities in the boundary layer,
$ {\textit{Re}}_{{cr}}$
, is
$ 4.5 \times 10^4$
. As the maximum
$ {\textit{Re}}_{{s}}$
investigated remains below this threshold, the boundary layer surrounding the droplet remains laminar throughout the experimental conditions. Understanding the airflow is crucial, as a strongly turbulent boundary layer, as highlighted by Stumpf et al. (Reference Stumpf, Qezeljeh, Kamal, Dezitter, Martuffo, Roisman and Hussong2024), can lead to droplet deformation and subsequent aerodynamic rebound. The disc is composed of an infrared-transmissive material (aluminium oxide,
$\alpha$
-
$\text{Al}_2\text{O}_3$
, also known as sapphire). To ensure accurate infrared measurements, the impact surface of the disc is coated with a highly emissive graphite-based paint. The influence of this coating on the measurements is considered negligible, an assumption widely supported in the literature (Chaze et al. Reference Chaze, Caballina, Castanet, Pierson, Lemoine and Maillet2019; Gholijani et al. Reference Gholijani, Schlawitschek, Gambaryan-Roisman and Stephan2020; Schmidt et al. Reference Schmidt, Breitenbach, Roisman and Tropea2020). The estimated time required for the thermal boundary layer within the coating to develop fully is given by
$t \propto \epsilon ^2/\alpha _{{c}}$
, where
$\epsilon$
denotes the thickness of the coating, and
$\alpha _{{c}}$
is its thermal diffusivity. In these experiments, the average thickness of the coating is
$ 8 \, \unicode{x03BC} \textrm{m}$
, and the thermal diffusivity is approximately
$ 2 \times 10^{-6} \, \textrm{m}^2\,\textrm{s}^{- 1}$
. This yields an estimated development time of
$t \approx 0.032$
ms, which is significantly shorter than the temporal resolution of
$\varDelta t = 0.15$
ms given by the infrared camera. The infrared camera is positioned beneath the disc to capture transient temperature measurements during droplet impact. In the experiments, droplet temperature, diameter and substrate speed are systematically varied. The substrate temperature, monitored with temperature sensors, varied with ambient room conditions and ranged from
$ 24.1$
–
$ 33.0 \,^\circ \textrm{C}$
across all experiments.
Table 1 summarises the experimental conditions tested in this study. A full factorial design was used to investigate the interactions between these parameters, with the droplet’s normal impact velocity held constant.
Table 1. Summary of the experimental conditions tested in this study. A full factorial design was used to systematically vary droplet diameter (
$D$
), substrate speed (
$U_{{d,t}}$
) and droplet temperature (
$T_{{d,0}}$
). The droplet’s normal impact velocity (
$U_{{d,n}}$
) was held constant, while key non-dimensional parameters, including the normal and tangential Reynolds (
${\textit{Re}}_{{d,n}}$
,
${\textit{Re}}_{{d,t}}$
) and Weber numbers (
${\textit{We}}_{{d,n}}$
,
${\textit{We}}_{{d,t}}$
), were computed over the range of experimental conditions.

2.2. Calibration of the infrared camera
The infrared camera is calibrated in situ using the set-up illustrated in figure 3. A copper plate, cooled by a chiller, extends beyond the camera’s entire field of view (FOV), ensuring uniform calibration across the measurement area.

Figure 3. Schematic of the calibration set-up for the infrared camera. The set-up includes a temperature-controlled copper plate thermally coupled to the sapphire disc via a thin layer of thermal paste, ensuring uniform heat transfer. A thermocouple is embedded at the surface to monitor and maintain the substrate temperature accurately. The sapphire disc, coated with a high-emissivity material, serves as the substrate for infrared measurements. The infrared camera captures the temperature distribution at the coating’s bottom surface within its FOV.
Temperature images are not directly obtained from the infrared camera; instead, the digital level (
$DL$
) recorded by the camera is converted into temperature values. The calibration process involves incrementally increasing the temperature of a reference surface from
$16$
–
$35\,^\circ \textrm{C}$
, with calibration images captured at
$1\,^\circ \textrm{C}$
intervals. The variation in
$DL$
is recorded as a function of the surface temperature (
$T$
), establishing a calibration curve. This relationship between
$DL$
and
$T$
is governed by Planck’s law, which describes the spectral radiant emittance (
$W_{b}$
) of an object as a function of its temperature,
\begin{align} W_{{b}}(\lambda , T_{\textit{obj}}) = \frac {2 \pi h c^2}{\lambda ^5} \frac {1}{\left (e^{\frac {h c}{\lambda k T_{\textit{obj}}}} - 1\right )}, \end{align}
where
$h$
is Planck’s constant,
$k$
is Boltzmann’s constant,
$c$
is the speed of light,
$\lambda$
is the wavelength and
$T_{\textit{obj}}$
is the temperature of the object.
Numerous studies, including those by Martiny et al. (Reference Martiny, Schiele, Gritsch, Schulz and Wittig1996), Schulz (Reference Schulz2000), Ochs et al. (Reference Ochs, Horbach, Schulz, Koch and Bauer2009), Elfner et al. (Reference Elfner, Schulz, Bauer and Lehmann2017), Falsetti, Sisti & Beard (Reference Falsetti, Sisti and Beard2021) and Sisti et al. (Reference Sisti, Falsetti, Beard and Chana2021), have based their infrared camera calibration on the above equation. By rearranging the terms and isolating
$T$
, the following semiempirical relation is obtained:
\begin{align} T = \frac {r}{\ln \left (\frac {b}{I_{\textit{obj}}} + f\right )}, \end{align}
where
$I_{\textit{obj}}$
is the digital level recorded by the camera, and
$r$
,
$b$
and
$f$
are calibration parameters. These parameters were estimated in situ using a thermocouple as the reference and a nonlinear least-squares fitting procedure applied to the calibration images obtained with the set-up illustrated in figure 3.
From the calibration process, two types of errors are quantified: the temporal noise in the temperature readings and the corresponding noise-equivalent heat flux. The temporal noise,
$\sigma _{{T}}$
, is defined as
\begin{equation} \sigma _{{T}} = \frac {1}{N_{\textit{pixels}}} \sum _{j=1}^{N_{\textit{pixels}}} \sigma _{{T,j}}, \end{equation}
where
$\sigma _{{T,j}}$
represents the standard deviation of temperature fluctuations for pixel
$j$
across multiple frames
$i$
, given by
\begin{equation} \sigma _{{T,j}} = \sqrt {\frac {1}{N_{\textit{frames}}} \sum _{i=1}^{N_{\textit{frames}}} \left (T_{{i,j}} - \bar {T}_{{j}}\right )^2}. \end{equation}
Here,
$N_{\textit{pixels}}$
is the total number of pixels in the region of interest, and
$N_{\textit{frames}}$
is the number of frames captured during the calibration measurement. Here
$T_{{i,j}}$
denotes the temperature of pixel
$j$
in frame
$i$
, and
$\bar {T}_{{j}}$
is the average temperature of pixel
$j$
across all frames. The overall temperature noise,
$\sigma _{{T}}$
, is computed as the average of
$\sigma _{{T,j}}$
over all pixels within the region of interest, assuming a spatially uniform temperature distribution.
The calculated temporal noise,
$\sigma _{{T}}$
, is
$0.21\,^\circ \textrm{C}$
. This value quantifies the variation in temperature readings over time. The corresponding noise-equivalent heat flux is determined separately through numerical simulation of the transient Fourier heat equation applied to the substrate. By analysing the artificial heat flux generated by temperature fluctuations between consecutive frames, the heat flux noise is estimated to be
$10^4 \, \textrm{W m}^{- 2}$
. Further details on the heat flux calculations are provided in § 2.4.
2.3. Temperature measurements
The experimental contact temperature values were directly obtained from the infrared camera measurements. A time-lapse sequence of the thermal evolution during droplet impact is shown in figure 4. The colder region represents the area wetted by the droplet, delineated by a sharp temperature gradient in the vicinity of the three-phase contact line. The contact area between the droplet and the substrate increases over time. Figure 4(a–d) correspond to four successive time steps: 0.57 ms, 1.03 ms, 1.61 ms and 2.07 ms after impact. As time progresses, the expansion of the contact area results from both droplet spreading and entrainment due to the relative motion with the substrate. To determine a single representative value for the contact temperature,
$T_{{c}}$
, the average temperature over the droplet–surface contact area was used. This approach is justified by the uniformity of the temperature across the contact region, as evident in figure 4. A slight deviation, specifically a lower temperature difference of approximately
$0.2 {-} 0.3 \,^\circ \textrm{C}$
, is observed in the initial contact area of the droplet, as evidenced in figure 4(d). This phenomenon can be attributed to the influence of the droplet’s initial normal velocity, which reduces the contact resistance (Aziz & Chandra Reference Aziz and Chandra2000) and consequently results in a slightly lower contact temperature in this region.
The contact temperature evolutions derived from experiments are given as a function of time for four different substrate velocities in figure 5(a–d). In addition to the experimental values and their associated errors, we compare our model with the models of Seki et al. (Reference Seki, Kawamura and Sanokawa1978) and Roisman (Reference Roisman2010), (1.2) and (1.3), respectively. These models predict a constant contact temperature, marked in figure 5 by dashed lines. At the early stages of impact, experimental results show a different behaviour, with the contact temperature decreasing monotonically before reaching the predicted values. We hypothesise that the thermal resistance at the substrate surface causes the temperature evolution. In this study, an improved model has been proposed, which accounts for both convective effects and thermal resistance at the interface. Exemplary results of this model are plotted as blue curves in figure 5(a–d). The theoretical framework for the current model is detailed in § 3.

Figure 4. Time-lapse sequence showing the thermal evolution during the impact of a droplet on a moving substrate, recorded using an infrared camera. The droplet, with a diameter of 1.95 mm and an initial temperature of 0.5
$\,^\circ$
C, impacts a substrate rotating at 800 r.p.m. (6.12 m s−1). Snapshots are taken at different times following impact: 0.57 ms, 1.03 ms, 1.61 ms and 2.07 ms. The temperature distribution within the droplet-substrate contact region remains mostly uniform, with slight deviations observed at the leading edge during initial contact.

Figure 5. Contact temperature evolution (
$T_{{c}}$
) over time at different rotational speeds (
$n$
). The panels correspond to (a)
$n = 200\,\text{r.p.m.}$
, (b)
$n = 400 \, \text{r.p.m.}$
, (c)
$n = 600 \, \text{r.p.m.}$
and (d)
$n = 800 \, \text{r.p.m.}$
, for a droplet diameter of
$D = 2.45 \, \text{mm}$
and an initial droplet temperature of
$T_{d,0} = 0.5 \, \,^\circ \text{C}$
. The experimental data (red curve) is compared with the theoretical predictions of the current model (blue curve), with the shaded area indicating the measurement error of the experimental data.
2.4. Heat flux evaluation
Heat flux at the wall cannot be directly retrieved from the infrared camera; therefore, a numerical simulation is performed based on the contact temperature measured by the camera. The Fourier heat diffusion equation is solved within the substrate as follows:
where
$\alpha _{{w}}$
is the thermal diffusivity of the substrate, and the velocity vector
$\boldsymbol{w}$
represents the substrate’s motion, as the measurements are taken in the laboratory reference frame. Thermal properties are assumed to remain constant throughout the analysis.
The computational mesh is designed such that the top layer emulates the camera pixels in a one-to-one mapping, with each pixel assigned a velocity vector derived from calibration images acquired during the motion-calibration procedure. The mesh spacing through the substrate thickness and the numerical time step are
$3.8\,\unicode{x03BC} \text{m}$
and
$30.16\,\unicode{x03BC} \text{s}$
, respectively. Grid- and time-independence studies are reported in Appendix A. At the interface, a Dirichlet boundary condition is applied, where the temperature is specified using data from the infrared camera. Along the sides of the domain, zero-gradient boundary conditions are imposed. At the bottom, a zero-gradient condition is also applied. This choice is based on the consideration that the substrate is assumed to remain at room temperature, and the heat flux at the bottom is expected to be negligible.
The validity of the assumption is assessed by analysing the effect of the entrained air within the boundary layer, which can lead to heating and a slight increase in temperature. To estimate this effect, the problem is treated analytically by solving a system of ordinary differential equations (ODEs) as described by Millsaps & Pohlhausen (Reference Millsaps and Pohlhausen1952), with the boundary condition modified to impose an adiabatic wall instead of a constant temperature. The details of the scaled temperature solution are provided in Appendix B.
Under these conditions, the maximum expected temperature increase for the highest substrate speed is
where
$S(0)$
and
$Q(0)$
are reduced parameters evaluated at the wall based on the ODE solution.
Aside from the theoretical description, this statement is further supported by the experimental findings of Cardone, Astarita & Carlomagno (Reference Cardone, Astarita and Carlomagno1994). While the Nusselt number
$\textit{Nu}$
reported in Cardone et al. (Reference Cardone, Astarita and Carlomagno1994) could, in principle, be used to calculate the heat transfer coefficient
$h_{\textit{tc}}$
for implementing a Robin-type boundary condition, the combination of a very small temperature difference and the uncertainty associated with the coefficient led us to adopt a zero-gradient boundary condition instead. Moreover, it is well established that for elliptic and parabolic equations, boundary effects diminish rapidly with increasing distance from the boundary (Evans Reference Evans2010).
Finally, preliminary simulations with both bottom boundary conditions (adiabatic and Robin) show that the heat flux at the impact-side interface (
$z=0$
) is unchanged. Any differences occur only near the bottom of the substrate, where the magnitude of
$\varDelta q = q_{{Robin}}-q_{{Neumann}}$
is one to two orders of magnitude below the experimental noise, and thus irrelevant for the reconstructed interfacial flux.
3. Heat transfer in a spreading drop, accounting for the thermal contact resistance
3.1. Problem formulation
The flow and temperature fields in the spreading droplet must satisfy the continuity, momentum and energy equations (Bird, Stewart & Lightfoot Reference Bird, Stewart and Lightfoot1966). We assume incompressible flow, as the time scale and spatial region affected by compressibility are negligible in our experiments. This assumption has been thoroughly discussed and validated by Weiss & Yarin (Reference Weiss and Yarin1999). Under these conditions, the governing mass balance, momentum balance and energy balance equations for the Newtonian flow
$\boldsymbol{\upsilon }$
, pressure
$p$
and temperature
$T$
are written as
The decoupling of the hydrodynamic problem from the thermal problem is a direct result of assuming constant thermophysical properties. This simplifies the analysis, allowing the temperature field to be solved independently of the velocity and pressure fields. This assumption is appropriate for the regime considered: temperature changes in the liquid remain moderate and are confined to a thin layer near the interface, variations of
$\rho$
,
$c_p$
and
$k$
are much smaller than the change in viscosity, which, in turn influences the thermal problem only indirectly via the advective correction to the similarity velocity
$w(z,t)$
. A more detailed validation and sensitivity analysis is provided in Appendix C. However, while valid in our parameter range, at larger temperature gradients (leading to more pronounced property variations), at higher shear rates where viscous heating is non-negligible, or when thermocapillary effects are sufficiently strong to modify the velocity field, a more thorough assessment would be required.

Figure 6. Illustration of droplet impact on a rough substrate, highlighting the interface between the droplet and the substrate. The zoomed-in view emphasises the roughness of the substrate and the liquid–substrate contact area, with the white regions representing entrapped air. The red arrows in the zoomed area represent the heat flux, illustrating local variations where regions with better contact exhibit higher heat flux, while areas with entrapped air show reduced heat flux.
Consider now the Cartesian coordinate system
$\{x,y,z\}$
with the unit base vectors
$\{\boldsymbol{e}_x, \boldsymbol{e}_y, \boldsymbol{e}_z\}$
with
$\boldsymbol{e}_z$
normal to the substrate interface directed towards the drop region. The velocity field in the fluid, which also accounts for the viscous effects, is defined as
$\boldsymbol{\upsilon } = u\boldsymbol{e}_x + \upsilon \boldsymbol{e}_y + w\boldsymbol{e}_z$
. The initial conditions and the boundary conditions far from the interface
$z=0$
are expressed as
where
$\boldsymbol{\upsilon }_0$
is the velocity in the outer region of the drop far from the viscous boundary layer. The expression for the velocity
$\boldsymbol{\upsilon }_0$
is obtained as the solution for the inviscid flow in the drop (Yarin & Weiss Reference Yarin and Weiss1995; Roisman Reference Roisman2010). Here,
$T_{{d,0}}$
and
$T_{{w,0}}$
denote the initial temperatures of the droplet and the substrate, respectively.
The no-slip and zero-flux boundary conditions yield the following boundary conditions for the flow at the interface:
In this study, the temperature discontinuity arising from the presence of contact resistance between the substrate and the spreading droplet is taken into account. The physical meaning of this thermal resistance is illustrated in figure 6. The thermal effects are influenced by the morphology of the substrate and air entrapment at the interface. The region in the substrate and in the drop, associated with the thermal resistance, is characterised by an effective thickness
$h_{{R}}$
and effective heat conductivity
$k_{{R}}$
. In this case the heat flux can be estimated in the form
$q \approx k_{{R}} (T_{{d}}-T_{{w}})/h_{{R}}$
, where
$h_{{R}}/k_{{R}} \equiv R_{{c}}$
is the thermal contact resistance as already proposed by Kapitza (Reference Kapitza1941), Swartz & Pohl (Reference Swartz and Pohl1989).
The thermal boundary conditions at the interface are expressed as
The first equation in (3.4) represents the energy balance at the surface, neglecting the heat accumulated due to the heating of the microscopic contact region, and the second term represents the temperature jump at this contact region.
In this study, we treat
$R_c$
as a single, casewise constant effective value, i.e. a lumped, area-averaged interfacial impedance over the inner contact zone. As microroughness and lamella pressure create heterogeneous microcontacts, the interfacial impedance is a function of time and space
$(R_c = R_c(x,y,t))$
; however, within our 1-D framework, only the area-averaged resistance enters the boundary condition. This simplification is supported by spatially and temporally smooth infrared temperature fields in the central patch of our wetted area (figure 4). We explicitly restrict model–data comparisons with the footprint interior, away from the lamella and rim where a constant
$R_c$
assumption would be violated. The same constant
$R_c$
closure is common in related droplet-impact and thermal-spray analyses to capture interfacial cooling processes (Dhiman & Chandra Reference Dhiman and Chandra2005; Nastic, Pershin & Mostaghimi Reference Nastic, Pershin and Mostaghimi2024). Using this boundary condition, rather than enforcing thermal continuity, reflects the physical reality of many practical systems, where contact resistance cannot be neglected, and highlights a key distinction from prior studies of drop-based cooling that assume perfect thermal continuity at the interface.
3.2. Viscous flow in the drop region
The similarity solution for flow within a lamella, expressed in Cartesian coordinates system
$\{x,y,z\}$
, fixed at the substrate surface, has been obtained in Roisman (Reference Roisman2009, Reference Roisman2010),
Here,
$f(\xi )$
and
$g(\xi )$
are, respectively, the dimensionless similarity profiles for the tangential and normal velocity components of the boundary-layer flow, where
$\xi$
is the similarity variable. At large values of
$\xi$
, this field approaches the outer inviscid flow in the spreading drop. The outer solution (Yarin & Weiss Reference Yarin and Weiss1995; Roisman Reference Roisman2010), which satisfies exactly the mass and momentum balance equations is
where
$U_x$
is the tangential,
$x$
-component of the impact velocity relative to the substrate. Here,
$\tau$
is a time constant; for the present initial condition it takes the value
$\tau \approx 0.25\,D_0/U_0$
(see Roisman et al. Reference Roisman, Berberović and Tropea2009).
Substituting the expression (3.5) in the mass and momentum balance (3.1a ) and (3.1b ) yields the following system of ODEs:
which can be solved computationally, subject to the no-penetrability conditions and the no-slip conditions at the wall and continuity of the velocity in the outer region
For completeness, the functions
$g(\xi )$
and
$f(\xi )$
first reported in Roisman (Reference Roisman2010) are plotted in figure 7.

Figure 7. Numerical solutions for the scaled velocity components as functions of the similarity variable
$\xi$
. Panel (a) shows the scaled vertical velocity
$g(\xi )$
, while panel (b) presents the scaled in-plane velocity
$f(\xi )$
, representing the velocity components in the
$x$
–
$y$
plane. These results are derived from the coupled ODEs described in (3.7a
), (a) The scaled vertical velocity component
$g(\xi )$
as a function of
$\xi$
. (b)The scaled in-plane velocity component
$f(\xi )$
(in the
$x$
–
$y$
plane) as a function of
$\xi$
.
3.3. Heat transfer in the drop and substrate at high Prandtl numbers
${\textit{Pr}}_{d}\gg 1$
Assuming contact resistance, the temperature fields in the spreading drop
$z\geqslant 0$
and the solid substrate
$z\leqslant 0$
are represented using the known analytical solution for heat conduction. For the solid region, the solution that exactly satisfies the energy equation together with the boundary and initial conditions is given in Carslaw & Jaeger (Reference Carslaw and Jaeger1959) as
where
$T_{c0}$
is a constant temperature equal to the asymptotic value of the contact temperature at long times, and
$t_R$
is a characteristic time associated with the duration of the propagation of the thermal boundary layer inside the contact region of thickness
$h_R$
. These constants will be determined from the boundary conditions. Here
$\zeta _w$
is the similarity variable in the solid region. The energy equation in the liquid region has to account for the heat convection. Therefore, in contrast to the purely diffusive semi-infinite solution used for the solid (Carslaw & Jaeger Reference Carslaw and Jaeger1959), we adopt a similarity ansatz that extends that diffusion form to include advection. We thus write for the spreading drop
where
$\alpha _{{d}}$
is the thermal diffusivity of the drop,
$c_d$
is a constant and the function
$\theta (\zeta _d)$
is the self-similar component of the temperature field associated with convective transport in the drop. Here
$\zeta _d$
is the similarity variable in the liquid region. In essence, the exponential factor in (3.9a
) encodes the effects of finite contact resistance, and the classical
$\operatorname {erf}(\zeta )$
term describes the response of a semi-infinite conductor under pure thermal diffusion. In the droplet, this term is replaced by the self-similar convective kernel
$\theta$
to account for thermal diffusion and advection in the flow. The kernel
$\theta$
is obtained from early-time asymptotics of the coupled advection–diffusion problem, as described later.
Expression (3.9a ) satisfies the heat conduction equation in the wall exactly. However, the form (3.9c ) yields only an approximate solution for the temperature field in the drop if the convective terms are smaller than the conductive terms. In the case of a large Prandtl number,
the thickness of the thermal boundary layer is much smaller than that of the viscous boundary layer. In this case, since both the value of the normal velocity and its gradient at the wall are equal to zero, the convection effects are indeed small. The validity of this assumption is assessed in Appendix D, where we compare the closed-form model with an independent advection–diffusion numerical reference. The asymptotic expression for the gradient of the drop temperature taking into account a small value of
$t_R/t$
is obtained from (3.9c
) in the form
This approximate expression can be used for the estimation of the small convective terms in the energy equation if the Prandtl number is large.

Figure 8. Theoretically predicted values of
$\theta (\zeta _d)$
at various values of the Prandtl number.
Substitution of expressions (3.9) and (3.11) in the energy (3.1c ) yields the following similarity equation:
\begin{equation} \theta ''(\zeta _d) + 2 \zeta _d \theta '(\zeta _d) + 4 {\textit{Pr}}_{{d}}^{1/2} g\left [\frac {\zeta _d }{{\textit{Pr}}_{{d}}^{1/2}}\right ] \left \{\theta '(\zeta _d) + \exp \left(-\zeta _d^2\right)\theta '(0)\right \}=0, \end{equation}
where the dimensionless function
$g$
is determined from the similarity solution for the drop velocity field.
This ODE has to be solved subject to the boundary conditions

Figure 9. Theoretically predicted values of
$\theta '(0)$
and
$c_{{d}}$
as a function of the Prandtl number.

Figure 10. Comparison of experimental contact temperatures with theoretical predictions for various models: (a) Seki et al. (Reference Seki, Kawamura and Sanokawa1978); (b) Roisman (Reference Roisman2010); (c) the current model. Each plot includes experimental data (circles), a line of perfect agreement (black dashed) and experimental error bounds (red dashed).

Figure 11. Heat transfer coefficient,
$h_{\textit{tc}}$
, as a function of time,
$t$
, for different rotational speeds and droplet diameters. Experimental data (solid lines) are compared with theoretical predictions (dashed lines). Panels (a) and (b) correspond to substrate rotational speeds of 200 and 800 r.p.m., respectively. Shaded regions indicate measurement uncertainty.
In figure 8, the theoretical predictions for the dimensionless function
$\theta (\zeta _d)$
are shown for various Prandtl numbers. The computed values of
$\theta '(0)$
as a function of
${\textit{Pr}}_{{d}}$
are shown in figure 9. The value of the constants
$T_{c0}$
and
$t_R$
are determined from the boundary conditions (3.4) associated with the thermal contact resistance
where
$\theta (\zeta _d)$
is a dimensionless function given in (3.9c
),
$e_{{d}}$
and
$e_{{w}}$
are the thermal effusivities of the drop and substrate materials, respectively. Finally, the substrate temperature at the interface
$z=0$
and the heat flux are obtained using (3.9) in the form
Equations (3.15a
) and (3.15b
) incorporate the influence of substrate morphology, resulting in a time-dependent interfacial temperature. Unlike (1.2) and (1.3), they preserve both the conductive mechanisms of Seki et al. (Reference Seki, Kawamura and Sanokawa1978) and the convective effects described by Roisman (Reference Roisman2010). The value of the thermal contact resistance,
$R_{{c}}$
, required to determine the characteristic time scale
$t_{{R}}$
was obtained by fitting the analytical substrate–temperature evolution (3.15a
) to each experimental curve and subsequently averaging the results. The mean value was found to be
$2.39 \times 10^{-5}$
$\text{m} ^2{\rm K\, W^{- 1}}$
with a 95 % confidence interval of [
$1.73 \times 10^{-5}$
,
$3.65 \times 10^{-5}$
]
$\text{m} ^2{\rm K\, W^{- 1}}$
.
A comparison between the experimental results and the three models can be seen in figure 10(a–c). For the models of Seki et al. (Reference Seki, Kawamura and Sanokawa1978) and Roisman (Reference Roisman2010), which predict a time-independent interfacial temperature, we perform the comparison at the instant when the measured interfacial temperature attains its local minimum. The model of Seki et al. (Reference Seki, Kawamura and Sanokawa1978) has a rather fair agreement as shown in figure 10(a), which can be explained by the fact that for our investigated
${\textit{Pr}}_{{d}}$
, the convective effects are small and the underestimation of the contact temperature is balanced by the missing contact resistance condition. On the other hand, the model of Roisman (Reference Roisman2010) consistently underestimates the contact temperature due to the missing contact resistance term. The present model incorporates both heat convection and contact resistance, resulting in a very good agreement with the experimental data.
The heat transfer coefficient at the interface is readily derived by scaling the heat flux to the initial temperature difference between the substrate and the spreading drop as
where
$\varDelta T = T_{{d0}} - T_{{w0}}$
is the initial temperature difference between the droplet and the substrate. This quantity is plotted in figures 11(a) and 11(b) for the two extreme substrate velocities tested (
$\approx$
200 and 800 r.p.m.), each at varying droplet diameters. Intermediate velocities were also investigated and exhibited qualitatively similar trends; they are therefore omitted here for conciseness. As expected, neither the droplet diameter nor the substrate velocity significantly affects the heat flux. This is consistent with the theoretical model predictions. The results demonstrate good agreement between experimental data (solid lines) and theoretical predictions (dashed lines), as shown in figure 11, where shaded regions represent measurement uncertainty. The present model applies to early-time, conduction-dominated exchange with finite interfacial resistance and no phase change. Two practical bounds are worth noting: at very high substrate speeds, large shear rates can introduce viscous heating and strong advection, which are not included here; and for very small (micronscale) droplets, the semi-infinite-liquid and locally 1-D assumptions may fail and free-surface effects may become important. Conversely, for larger droplets, the assumptions remain valid. Furthermore, regimes involving phase change (boiling, evaporation, solidification) are outside the scope of this study. For completeness, we propagate the estimated uncertainty in the interfacial resistance
$R_c$
into the predicted heat-transfer coefficient
$h_{tc}(t)$
; the procedure and the resulting 95 % confidence intervals are given in Appendix E. These intervals are consistent with the experimental ones. While the contact–temperature points in figure 10 differ only modestly through the models, the corresponding heat flux, and therefore
$h_{tc}(t)$
, shows larger differences between the models. This arises from including the contact–resistance boundary condition, which yields a finite heat flux at
$t=0$
, in contrast to the initial singularity predicted by the models of Seki et al. (Reference Seki, Kawamura and Sanokawa1978) and Roisman (Reference Roisman2010). This finite behaviour highlights the importance of incorporating realistic boundary conditions in heat transfer models to accurately capture interfacial dynamics during droplet impact.
4. Conclusion
This study presents a comprehensive theoretical and experimental investigation of the thermal interactions during the impact of liquid droplets on a moving solid substrate. We develop a closed-form model that simultaneously accounts for conduction in the wall, convective heat transport within the spreading droplet and substrate-morphology effects via a finite interfacial contact resistance
$R_c$
. By extending the self-similar spreading solution of Roisman (Reference Roisman2010), the revised framework yields closed-form expressions for both interfacial temperature and heat flux that remain finite and physically realistic throughout the impact process. The model is validated through high-speed, spatially resolved infrared thermography, which captures the 2-D, transient temperature evolution at the droplet–substrate interface during impingement on a rotating, infrared-transmissive sapphire disc coated with a highly emissive graphite-based layer. The experimental temperature maps are used to numerically reconstruct local interfacial heat fluxes, which show strong agreement with the theoretical predictions. Across a range of droplet diameters, substrate velocities, and thermal conditions – including normal and tangential Reynolds numbers (
$\displaystyle {\textit{Re}}_{{d,n}}\approx 2660{-}4406$
,
$\displaystyle {\textit{Re}}_{{d,t}}\approx 1566{-}13\,608$
) and Weber numbers (
$\displaystyle {\textit{We}}_{{d,n}}\approx 200{-}292$
,
$\displaystyle {\textit{We}}_{{d,t}}\approx 90{-}1169$
), the thermal behaviour is found to be insensitive to variations in droplet size and substrate speed. This confirms the robustness of the model and supports the theoretical prediction that these parameters have a negligible influence on interfacial heat transfer under the studied conditions.
Supplementary movies
Supplementary movies are available at https://doi.org/10.1017/jfm.2025.10978.
Acknowledgements
The authors gratefully acknowledge Dr-Ing. M. Lausch for his invaluable insights and stimulating discussions throughout this work, which greatly enhanced its quality and rigour. We also thank O.U.R. Siddiqui for his assistance with the experimental programme. Parts of this work have been improved grammatically and stylistically using language models, including Chat-GPT (OpenAI) and DeepL Write (DeepL SE).
Funding
The work of R.K. was funded by the European Union’s Horizon Europe research and innovation programme under the Marie Skłodowska-Curie grant agreement no. 101072551 (TRACES). The work of J.H. and I.V.R. is partially supported by the joint DFG/FWF Collaborative Research Centre CREATOR (DFG: Project-ID 492661287/TRR 361; FWF: 10.55776/F90) at TU Darmstadt, TU Graz and JKU Linz.
Declaration of interests
The authors report no conflict of interest.
Appendix A. Grid and time independence study
The through-thickness mesh spacing and the time step used in the production runs are
$\Delta z = 3.8\,\unicode{x03BC} \text{m}$
and
$\Delta t = 30.16\,\unicode{x03BC} \text{s}$
, respectively. Since temperature converges with relatively high time and grid spacings, grid convergence is assessed by comparing the interfacial heat flux
$q(t)$
on successive refinements. We define the maximum relative error over the common time window
$I$
as
where
$i$
and
$i{+}1$
denote successive refinement levels, with
$i{+}1$
finer than
$i$
. We declare grid (and, analogously, time step) independence when
$E_\infty ^{(i)} \leqslant 2\,\%$
. The corresponding grid- and time-independence plots are shown in figures 12 and 13.

Figure 12. Surface heat flux
$q(t)$
for several through-thickness grids.

Figure 13. Surface heat flux
$q(t)$
for a fixed through-thickness grid (
$\varDelta _z=3.8\,\unicode{x03BC} \text{m}$
) at several time steps.
Appendix B. Airflow past a rotating disc
For the sake of clarity, we report here the solution of the three-dimensional (3-D) boundary layer problem of air due to a rotating disc. For the derivation of the equations, the interested reader is referred to Kármán (Reference Kármán1921) and Millsaps & Pohlhausen (Reference Millsaps and Pohlhausen1952).

Figure 14. Solutions to the 3-D boundary-layer problem for a rotating disc. (a) Hydrodynamic components
$H(\xi )$
,
$F(\xi )$
and
$G(\xi )$
. (b) Thermodynamic components
$S(\xi )$
and
$Q(\xi )$
.
A dimensionless system of ODEs can be found by imposing
For the thermal distribution,
The non-dimensional incompressible Navier–Stokes equations simplify to the following system:
where
subject to the boundary conditions
The corresponding energy equation reduces to
subject to the following boundary conditions:
Here,
${{\textit{Pr}}_a}$
represents the Prandtl number of air. The first set of four ODEs, given in (B3)–(B6), describes the hydrodynamic part of the problem, as proposed by Kármán (Reference Kármán1921) and solved numerically by Cochran (Reference Cochran1934). The solution to these equations is shown in figure 14(a), which presents the non-dimensional vertical, radial, and azimuthal velocity components
$H(\xi )$
,
$F(\xi )$
and
$G(\xi )$
, respectively. The second set, (B9), (B10), governs the thermodynamic problem. The solution, displayed in figure 14(b), represents the non-dimensional temperature distribution
$S(\xi )$
and
$Q(\xi )$
. The values of
$S(0)$
and
$Q(0)$
used in (2.6) are, respectively, 0.44 and 7.95.

Figure 15. Temperature rise in the droplet,
$\Delta T_d(z,t)=T_d-T_{d0}$
, at selected times over
$z$
.
Appendix C. Sensitivity to temperature-dependent properties
We assess the sensitivity of the analysis to temperature-dependent properties and, in particular, justify the constant-property approximation adopted in the main text. In the regime considered, interfacial heating is confined to a thin liquid layer adjacent to the wall (
$\approx$
50
$\unicode{x03BC} \textrm{m}$
; see figure 15), and variations of
$\rho$
,
$c_p$
and
$k$
are subdominant
$\bigl ({O}(5\,\%)\bigr )$
. The dominant temperature dependence enters through the viscosity, which affects the thermal problem only indirectly via the advective correction to the similarity velocity,
with
$\xi = z/\sqrt {\nu t}$
. Thus, the influence of
$\mu(T)$
is moderated both by the square-root dependence on
$\nu$
and by the fact that, over the time window of interest, advection remains weaker than conduction.
To test this, we solve a 1-D coupled wall–liquid problem and compare a baseline case (properties evaluated at
$T_{d0}$
) with a conservative ‘hot’ case in which the liquid properties are evaluated at
$T_{d0}+\varDelta T_{d,max }$
. The quantity
$\varDelta T_{d,max }$
is determined from a reference case similar to our experimental conditions, with the sapphire substrate at
$30\,^\circ \textrm{C}$
and the droplet at
$0\,^\circ \textrm{C}$
. Figure 15 shows the maximum temperature increase on the liquid side within the chosen time window, which is
$\approx 18\,^\circ \textrm{C}$
. Comparing the wall-side temperature profiles for the hot-property and baseline cases shows that the difference
$T_{{w}}^{{hot}}-T_{{w}}^{{base}}$
remains negligible over the examined times (figure 16). Even under this conservative comparison, these results support the use of temperature-independent properties at leading order for the conditions examined: the thermal perturbation is shallow and localised, conductive transport dominates over the advective correction, and viscosity-driven changes in the similarity velocity have a negligible effect on the predicted interfacial heat transfer.

Figure 16. Wall-side temperature difference profiles.
Appendix D. Large-
${\textit{Pr}}$
validity check
We assess the large-
${\textit{Pr}}$
closure by comparing analytical (dashed) and numerical (solid) wall-side temperature profiles, plotted in the non-dimensional form
for three fluids spanning orders of magnitude in
${\textit{Pr}}$
: glycerol (
${\textit{Pr}}\approx 1000$
), water (
${\textit{Pr}}\approx 7$
) and mercury (
${\textit{Pr}}\approx 0.01$
). The corresponding wall-side non-dimensional temperature profiles
$\Theta$
are shown in figure 17. For water and glycerol, the two curves agree closely over the window examined, supporting the use of the large-
${\textit{Pr}}$
closure; discrepancies increase for low-
${\textit{Pr}}$
mercury, as expected. We use a temperature difference between substrate and liquid of
$30\,^\circ \textrm{C}$
to match the experiments, with
$T_{d,0}=20\,^\circ \textrm{C}$
and
$T_{w,0}=50\,^\circ \textrm{C}$
. This choice, rather than
$0{-}30\,^\circ \textrm{C}$
, avoids invoking supercooling for glycerol.

Figure 17. Wall-side non-dimensional temperature profiles
$\varTheta$
at six times (0.10–5.00 ms).
Appendix E. Propagation of
$\boldsymbol{R}_{\boldsymbol{c}}$
uncertainty into
$\boldsymbol{h}_{\boldsymbol{tc}}$

Figure 18. Propagation of
$R_c$
uncertainty into
$h(t)$
: best-fit curve
$h(t;\hat R_c)$
with 95 % confidence band computed from (E13).
We quantify how the uncertainty of the fitted interfacial resistance
$R_c$
affects the modelled heat–transfer coefficient
$h_{tc}(t)=q(t)/\Delta T$
, where
Recall the closed–form expression used in the main text,
with
as defined in the convection closure (see § 3).
For compactness, set
so that
Let
$\hat R_c$
be the point estimate and
$\sigma _{R_c}$
its standard error (e.g. from a 95 % CI
$[R_c^{\textit{low}},R_c^{\textit{high}}]$
via
The analytic sensitivity of
$h_{tc}$
to
$R_c$
at
$\hat R_c$
is
A pointwise 95 % confidence band for
$h_{tc}$
due solely to the uncertainty in
$R_c$
is then































































