Hostname: page-component-68c7f8b79f-xmwfq Total loading time: 0 Render date: 2025-12-18T11:09:29.757Z Has data issue: false hasContentIssue false

Making a deposit: the role of substrate adsorption in coffee-ring formation in sessile evaporating droplets

Published online by Cambridge University Press:  18 December 2025

Madeleine R. Moore*
Affiliation:
Department of Mathematical Sciences, Loughborough University , Schofield Building, University Road, Loughborough LE11 3TU, UK
Hannah-May D’Ambrosio
Affiliation:
School of Mathematics & Statistics, University of Glasgow, University Place, Glasgow G12 8QQ, UK
Alexander W. Wray
Affiliation:
Department of Mathematics and Statistics, University of Strathclyde, Livingstone Tower, 26 Richmond Street, Glasgow G1 1XH, UK
*
Corresponding author: Madeleine R. Moore, m.r.moore@lboro.ac.uk

Abstract

A thin, evaporating sessile droplet with a pinned contact line containing inert particles is considered. In the limit in which the liquid flow decouples from the particle transport, we discuss the interplay between particle advection, diffusion and adsorption onto the solid substrate on which the droplet sits. We perform an asymptotic analysis in the physically relevant regime in which the Péclet number is large, i.e. ${\textit{Pe}}\gg 1$, so that advection dominates diffusion in the droplet except in a boundary layer near the contact line, and in which the ratio of the particle velocities due to substrate adsorption and diffusion is at most of order unity as ${\textit{Pe}}\rightarrow \infty$. We use the asymptotic model alongside numerical simulations to demonstrate that substrate adsorption leads to a different leading-order distribution of particle mass compared with cases with negligible substrate adsorption, with a significant reduction of the mass in the suspension – the nascent coffee ring reported in Moore et al. (J. Fluid Mech., vol. 920, 2021, A54). The redistribution leads to an extension of the validity of the dilute suspension assumption, albeit at the cost of breakdown due to the growth of the deposited layer, which are important considerations for future models that seek to accurately model the porous deposit regions.

Information

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

The evaporation of particle-laden droplets is ubiquitous, with a multitude of everyday applications in addition to industrial and engineering uses. The field is most famously concerned with the so-called ‘coffee-ring’ effect, whereby a spilled droplet of coffee leaves behind a stain that is darkest towards its edge – the eponymous ring (Deegan et al. Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten1997, Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000). In brief, the pinning of the contact line of the droplet on the substrate as it evaporates leads to an outward capillary flow to replenish the mass loss at the contact line, which carries particles with it, forming the ring.

This phenomenon is by no means restricted to coffee and may be exploited in situations where patterning is desired, such as in printing microcircuits or colloidal patterning (see e.g. Harris et al. Reference Harris, Hu, Conrad and Lewis2007; Mirbagheri & Hwang Reference Mirbagheri and Hwang2019; Zhou et al. Reference Zhou, Yu, Zou, Wang and Liu2019; Issakhani et al. Reference Issakhani, Jadidi, Farhadi and Bazargan2023). The pattern itself may also be analysed to reveal information in blood spatter analysis (Brutin & Starov Reference Brutin and Starov2018; Smith & Brutin Reference Smith and Brutin2018) and healthcare diagnostics (Smalyukh et al. Reference Smalyukh, Zribi, Butler, Lavrentovich and Wong2006; Sefiane, Duursma & Arif Reference Sefiane, Duursma and Arif2021). However, in other applications, the coffee-ring effect may be undesirable, particularly when uniform coatings are required in, for example, additive manufacturing of Q/OLED screens (Jiang et al. Reference Jiang, Zhong, Liu, He, Zou, Wang, Wang, Peng and Cao2016; Li et al. Reference Li, Duan, Shao, Zhang, Li, Huang and Yin2020).

With the wide variety of applications, it is perhaps unsurprising that the phenomenon remains at the centre of active research. There is a large body of literature concerning different aspects of the problem such as droplet evaporation (Wilson & D’Ambrosio Reference Wilson and D’Ambrosio2023), evaporation-driven internal flow (Gelderblom, Diddens & Marin Reference Gelderblom, Diddens and Marin2022) and techniques for controlling deposition (Mampallil & Eral Reference Mampallil and Eral2018). Of key interest is the role played by various physical effects in the particle deposition, particularly in the formation of the coffee ring, with studies highlighting the importance of substrate wettability (Nguyen, Hampton & Nguyen Reference Nguyen, Hampton and Nguyen2013; Patil et al. Reference Patil, Bange, Bhardwaj and Sharma2016), Marangoni effects (Kajiya & Doi Reference Kajiya2011; Abo Jabal et al. Reference Abo Jabal, Homede, Pismen, Haick and Leshansky2017; Malinowski et al. Reference Malinowski, Volpe, Parkin and Volpe2018), droplet geometry (Sáenz et al. Reference Sáenz, Wray, Che, Matar, Valluri, Kim and Sefiane2017; Wray & Moore Reference Wray and Moore2023, Reference Wray and Moore2024), gravity (Kim et al. Reference Kim, Gonçalves, Jung, Kim and Weon2021; Moore & Wray Reference Moore and Wray2023) and vapour shielding effects in the atmosphere due to masking (Harris et al. Reference Harris, Hu, Conrad and Lewis2007; D’Ambrosio et al. Reference D’Ambrosio, Wilson, Wray and Duffy2023) or the presence of other droplets in close proximity (Wray et al. Reference Wray, Wray, Duffy and Wilson2021; Wray & Moore Reference Wray and Moore2024).

Given that the coffee-ring effect occurs due to outward capillary flow, most studies regarding deposition, including the majority of those mentioned above, have focused on aspects of the problem that affect particle transport via the internal flow. In practice, however, in addition to this flow, the final deposit is also determined by the particle–particle, particle–free surface and particle–substrate interactions present. In recent years there has therefore been an emergence of analytical, numerical and experimental studies seeking to elucidate these effects (see e.g. Anyfantakis & Baigl Reference Anyfantakis and Baigl2015; Shao et al. Reference Shao, Duan, Hou and Zhong2020). In particular, it has been shown experimentally that the ring deposit is surprisingly robust and often persists (if in a somewhat reduced form) despite the enhancement of particle–particle interactions through, for example, increasing particle size and/or concentration (Malla, Bhardwaj & Neild Reference Malla, Bhardwaj and Neild2019) or of particle–substrate interactions through, for example, changing the acidity of the base fluid (Bhardwaj et al. Reference Bhardwaj, Fang, Somasundaran and Attinger2010), the addition of surfactant (Anyfantakis et al. Reference Anyfantakis, Geng, Morel, Rudiuk and Baigl2015) or modification of the particle and/or substrate charge (Yan et al. Reference Yan, Gao, Sharma, Chiang and Wong2008; Bridonneau et al. Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020). However, whilst a ring deposit often persists, the characteristics of the final deposit depend strongly on the electrostatic potential between the particles and the substrate (see e.g. Yan et al. Reference Yan, Gao, Sharma, Chiang and Wong2008; Bhardwaj et al. Reference Bhardwaj, Fang, Somasundaran and Attinger2010; Molchanov et al. Reference Molchanov, Roldughin, Chernova-Kharaeva and Senchikhin2019; Kumar, Basavaraj & Satapathy Reference Kumar, Basavaraj and Satapathy2023). Specifically, when the particles and the substrate are oppositely charged, the electrostatic force between them is attractive, resulting in more particles being deposited onto the substrate in the droplet bulk, i.e. a reduction of mass in the contact line deposit (see e.g. Anyfantakis & Baigl Reference Anyfantakis and Baigl2015; Devineau et al. Reference Devineau, Anyfantakis, Marichal, Kiger, Morel, Rudiuk and Baigl2016; Kumar et al. Reference Kumar, Basavaraj and Satapathy2023). We note that, despite the robustness of the ring deposit for typical spherical particles, studies on ellipsoidal particles have found that enhanced long-range particle–particle interactions can cause loosely packed clusters to aggregate at the free surface of the droplet, leading to more uniform deposits (Yunker et al. Reference Yunker, Still, Lohr and Yodh2011; Dugyala & Basavaraj Reference Dugyala and Basavaraj2014).

Previous numerical and analytical studies have examined how particle–substrate adsorption, hereafter referred to as substrate adsorption, may contribute to the formation or elimination of the coffee ring (see e.g. Widjaja & Harris Reference Widjaja and Harris2008; Siregar, Kuerten & Van der Geld Reference Siregar, Kuerten and Van der Geld2013; Zigelman & Manor Reference Zigelman and Manor2018a , Reference Zigelman and Manorb ; Sung, Wang & Harris Reference Sung, Wang and Harris2019; Erdem, Denner & Biancofiore Reference Erdem, Denner and Biancofiore2024; D’Ambrosio et al. Reference D’Ambrosio, Wray and Wilson2025b ). Specifically, Widjaja & Harris (Reference Widjaja and Harris2008) numerically investigated the effect of advective transport due to evaporation, colloidal diffusion and substrate adsorption on the deposition from a droplet undergoing diffusion-limited evaporation. They identified four distinct final deposit patterns, with a transition from a ring deposit when advective transport dominates, to a deposit concentrated at the droplet centre when substrate adsorption dominates. Extensions to this investigation have been proposed by including the effect of contact-line motion (Siregar Reference Siregar2012; Siregar et al. Reference Siregar, Kuerten and Van der Geld2013) and particle–substrate desorption (Sung et al. Reference Sung, Wang and Harris2019). More recently, Erdem et al. (Reference Erdem, Denner and Biancofiore2024) used the model proposed by Widjaja & Harris (Reference Widjaja and Harris2008) to consider the effect of colloidal diffusion, substrate adsorption, contact-line motion and thermal effects on the deposition of particles from a droplet evaporating according to the one-sided model (see e.g. Anderson & Davis Reference Anderson and Davis1995).

Zigelman & Manor (Reference Zigelman and Manor2018a ,Reference Zigelman and Manor b ) used numerical and analytical techniques to study the effect of substrate adsorption and particle–particle coagulation on the deposition from a thin sessile droplet undergoing spatially uniform evaporation under the assumption that colloidal diffusion is negligible. In particular, Zigelman & Manor (Reference Zigelman and Manor2018a ) identified a critical strength of substrate adsorption at which the concentration of particles within the droplet remains uniform throughout evaporation, and observed a transition from concentration profiles that are singular at the contact line below this threshold to profiles that are more concentrated at the centre of the droplet above this threshold, indicating a suppression of the coffee-ring effect. D’Ambrosio et al. (Reference D’Ambrosio, Wray and Wilson2025b ) extended the work of Zigelman & Manor (Reference Zigelman and Manor2018a ,Reference Zigelman and Manor b ) by investigating the concentration of particles adsorbed onto the substrate, as well as the evolution of masses within the bulk of the droplet, adsorbed onto the substrate, and in the ring deposit for a droplet undergoing either spatially uniform or diffusion-limited evaporation. In particular, they predict a transition from an adsorbed deposit more concentrated near the contact line of the droplet when radial advection due to evaporation dominates particle–substrate adsorption, to a deposit more concentrated near the centre of the droplet when particle–substrate adsorption dominates radial advection due to evaporation, and that the presence of substrate adsorption suppresses the formation of a ring deposit at the contact line for spatially uniform, but not for diffusion-limited, evaporation.

The aforementioned theoretical studies investigating substrate adsorption propose extensions to the basic model for the transport of colloid within a thin droplet originally described by Deegan et al. (Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000). In this model it is assumed that the colloid is sufficiently dilute that it has no effect on the flow within the droplet and that the appropriately defined Péclet number is sufficiently large that the radial transport is governed solely by advection. These assumptions are expected to break down in areas of high concentration, such as near the contact line of the droplet. In particular, close to the contact line, where the droplet thickness approaches zero, the particle concentration diverges resulting in a non-zero mass flux through the contact line (Deegan et al. Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000); this is usually interpreted as deposition of the particles into a ring.

This occurs even for evaporative models outside of diffusion-limited evaporation and will typically occur when there is radially outward flow (D’Ambrosio et al. Reference D’Ambrosio, Wilson, Wray and Duffy2023). While the inclusion of particle–substrate adsorption may be shown to reduce the mass transported to the ring, existing models in the diffusion-limited regime still exhibit this loss of mass (D’Ambrosio et al. Reference D’Ambrosio, Wray and Wilson2025b ).

Various authors have extended the analysis of Deegan et al. (Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000) in attempts to overcome the unphysical mass loss at the contact line by, for example, accounting for finite particle size effects, essentially allowing for particles to trap and/or jam close to the contact line and hence removing the singularity in the particle concentration profile (Ozawa, Nishitani & M Reference Ozawa and Nishitani2005; Popov Reference Popov2005; Eales et al. Reference Eales, Routh, Dartnell and Goddard2015; Kaplan & Mahadevan Reference Kaplan and Mahadevan2015; Coombs et al. Reference Coombs, Chubynsky and Sprittles2024a , Reference Coombs, Sprittles and Chubynskyb ). However, it has been shown that the assumptions of the dilute model fail before particle packing is expected to become significant and that there exists a boundary layer close to the contact line in which diffusion dominates particle transport (Moore et al. Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022; Moore & Wray Reference Moore and Wray2023). In particular, this allows the no-flux condition for the particle volume fraction, which is typically neglected in the reduction from a second-order to a first-order partial differential equation in the large-Péclet-number limit, to once again be imposed at the contact line.

The aim of the present work is therefore twofold. First, we extend the analysis of D’Ambrosio et al. (Reference D’Ambrosio, Wray and Wilson2025b ) to include particle diffusion and thereby address the loss of particle mass at the contact line observed therein. Second, we use both asymptotic and numerical analyses to analyse the dynamic distribution of particle mass between the adsorbed layer, the nascent coffee ring and the droplet bulk. These results are used to discuss different limitations of the dilute model. Where possible, we corroborate our findings with previous work in the literature.

The structure of the paper is as follows. In § 2, we formulate and derive the dimensionless model, before identifying the critical asymptotic regimes. The system is analysed asymptotically and numerically in the physically relevant limit of advection-dominated particle transport in § 3. The case ${\textit{Pe}}=O(1)$ is examined numerically in § 4. We give comparisons against suitable simulation and experimental results in § 5. We conclude with a summary and discussion in § 6.

2. Problem formulation

We consider a droplet of volatile fluid evaporating from a solid substrate as depicted schematically in figure 1. The droplet has initial volume $\tilde V_{\textit{init}}$ and circular contact radius $\tilde R$ . Here and hereafter, a tilde denotes a dimensional variable. The liquid has density $\tilde \rho$ , dynamic viscosity $\tilde \mu$ and surface tension coefficient $\tilde \sigma$ with the surrounding gas. Throughout this analysis, these are taken to be constant. The effect of gravity is assumed to be negligible so that the Bond number $\textit{Bo} = \tilde \rho \tilde g \tilde R^2/\tilde \sigma \ll 1$ , where $\tilde g$ is the constant acceleration due to gravity.

Figure 1. A droplet of volatile fluid lying on a solid substrate evaporates into the surrounding gas. The droplet contains inert particles that are transported by the competing effects of advection (yellow), diffusion (blue) and substrate adsorption (magenta). Advection to the pinned contact line is driven by the flow due to the evaporation. This increases the local particle volume fraction, driving a diffusive flux of particles near the contact line. Finally, particles may be arrested on the substrate due to adsorption.

The droplet evaporates into the surrounding gas, and is taken to be sufficiently small that evaporation is diffusion-limited and quasi-steady (Wilson & D’Ambrosio Reference Wilson and D’Ambrosio2023). We note that there are other common evaporation models applicable for different substrate–liquid–gas combinations, including the one-sided or kinetic evaporation model (Murisic & Kondic Reference Murisic and Kondic2011), but we choose to focus on the most commonly used diffusion-limited model here. Throughout the evaporation, we assume that the gas is quiescent and that the droplet contact line remains pinned on the substrate. Pinning, typically due to surface inhomogeneities, is crucial to the coffee-ring effect and is observed in practice for a wide range of liquids for the majority of the drying time (Deegan et al. Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten1997; Hu & Larson Reference Hu and Larson2002; Kajiya, Kaneko & M Reference Kajiya and Kaneko2008; Howard et al. Reference Howard, Archer, Sibley, Southee and Wijayantha2023). Moreover, pinning may be further amplified by particle accumulation at the contact line, extending the period of validity for the assumption (Orejon, Sefiane & Shanahan Reference Orejon, Sefiane and Shanahan2011; Weon & Je Reference Weon and Je2013; Larson Reference Larson2014). We note here that mechanisms for reducing or inhibiting pinning are used to diminish or eliminate the coffee ring, for example by exploiting weak contact angle hysteresis on hydrophobic substrates (Li, Sheng & Tsao Reference Li, Sheng and Tsao2013) or electrowetting (Eral et al. Reference Eral, Mampallil Augustine, Duits and Mugele2011).

We take cylindrical polar coordinates $(\tilde r, \theta , \tilde z)$ centred at the stagnation point within the droplet with the substrate lying in the plane $\tilde z = 0$ . For the purposes of this analysis, we assume that the droplet is axisymmetric, so that the resulting flow and transport problems are independent of the polar angle $\theta$ . Therefore, the liquid–vapour interface is denoted by $\tilde z = \tilde h (\tilde r, \tilde t)$ , while the pinned circular contact line is $\tilde r = \tilde R$ , $\tilde z = 0$ .

Throughout the modelling, we assume the particles are sufficiently dilute that the flow and transport equations decouple. This assumption will undoubtedly break down at later times when the particle volume fraction approaches the packing fraction (e.g. ${\approx}64\,\%$ for closely packed spheres), whereby finite particle size effects become important (see e.g. Popov (Reference Popov2005), Kaplan & Mahadevan (Reference Kaplan and Mahadevan2015), Moore, Vella & Oliver (Reference Moore, Vella and Oliver2021) and Coombs et al. (Reference Coombs, Chubynsky and Sprittles2024a , Reference Coombs, Sprittles and Chubynskyb ) for detailed discussions on the onset of finite particle size effects and potential modelling frameworks). We assess the possibility of breakdown due to finite particle size effects in further detail in § 3.5. Under the dilute assumption, the equations governing the motion of the flow are the Stokes equations,

(2.1) \begin{equation} \boldsymbol{\nabla }\boldsymbol{\cdot }\tilde {\boldsymbol{u}} = 0, \quad \boldsymbol{\nabla}\!\tilde{p}=\tilde \mu {\nabla} ^2 \tilde {\boldsymbol{u}} \quad \mbox{for} \quad \tilde r\lt \tilde R, \; 0\lt \tilde z\lt \tilde h(\tilde r, \tilde t), \end{equation}

where, since the flow is axisymmetric, the liquid velocity is $\tilde {\boldsymbol{u}} = \tilde {u}(\tilde r,\tilde z,\tilde t)\boldsymbol{e}_r + \tilde {w}(\tilde r,\tilde z, \tilde t)\boldsymbol{e}_z$ and $\tilde p(\tilde r, \tilde z, \tilde t)$ is the liquid pressure.

The Stokes equations must be solved subject to the usual conditions of no-slip and no-penetration on the substrate:

(2.2) \begin{equation} \tilde {\boldsymbol{u}}(\tilde r,0, \tilde t) = \boldsymbol{0} \quad \mbox{for} \quad \tilde r \lt \tilde R; \end{equation}

while on the liquid–air free surface, the stress conditions are given by

(2.3) \begin{equation} \boldsymbol{n}\boldsymbol{\cdot }\big(\!-\tilde p_{{\textit{atm}}}\boldsymbol{I} - \tilde {\boldsymbol{T}}\big) = -\tilde{\sigma}\tilde{\kappa}\boldsymbol{n} \quad \mbox{on} \quad \tilde z = \tilde h(\tilde r,\tilde t), \; \tilde r \lt \tilde R, \end{equation}

where $\boldsymbol{n}$ is an outward-pointing unit normal to the free surface, $\tilde \kappa = -\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{n}$ is the free-surface curvature, $\boldsymbol{I}$ is the identity tensor, $\tilde {\boldsymbol{T}}$ is the Newtonian stress tensor and $\tilde {p}_{{\textit{atm}}}$ is the constant atmospheric pressure; and finally the kinematic condition gives

(2.4) \begin{equation} \frac {\partial \tilde h}{\partial \tilde t}+\tilde u\frac {\partial \tilde h}{\partial \tilde r}-\tilde w=-\frac {\tilde J}{\tilde \rho } \quad \mbox{on} \quad \tilde z = \tilde h(\tilde r,\tilde t), \; \tilde r \lt \tilde R, \end{equation}

where $\tilde J/\tilde \rho$ represents the flux of liquid lost due to evaporation.

The volume fraction of the inert particles contained in the droplet, $\tilde \phi =\tilde \phi (\tilde r, \tilde z, \tilde t)$ , is governed by the advection–diffusion equation:

(2.5) \begin{equation} \frac {\partial \tilde \phi }{\partial \tilde t}+\tilde {\boldsymbol{u}}\boldsymbol{\cdot }\boldsymbol{\nabla }\tilde \phi =\tilde {D}_{{s}} {\nabla} ^2 \tilde \phi \quad \mbox{for} \quad \tilde r\lt \tilde R, \; 0\lt \tilde z\lt \tilde h(\tilde r, \tilde t), \end{equation}

where $\tilde {D}_{{s}}$ is the diffusivity of the particles in the liquid, which we take to be constant.

Accounting for motion due to evaporation, the no-penetration condition at the liquid–air interface is

(2.6) \begin{equation} \tilde {D}_{{s}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\tilde \phi =\frac {\tilde \phi \tilde J}{\tilde \rho } \quad \mbox{on} \quad \tilde z = \tilde h(\tilde r,\tilde t), \; \tilde r \lt \tilde R. \end{equation}

Similarly, to account for the loss of particles to adsorption on the substrate, we follow previous models on substrate adsorption by setting

(2.7) \begin{equation} \tilde {D}_{{s}} \frac {\partial \tilde \phi }{\partial \tilde z}=\tilde {k}_{{a}} \tilde \phi \quad \mbox{on} \quad \tilde {z} = 0, \end{equation}

where $\tilde {k}_{{a}}$ is the (constant) substrate adsorption velocity (Kurrat et al. Reference Kurrat, Ramsden and Prenosil1994, Reference Kurrat, Prenosil and Ramsden1997; Siregar Reference Siregar2012; Siregar et al. Reference Siregar, Kuerten and Van der Geld2013; Zigelman & Manor Reference Zigelman and Manor2018a ). Here it is assumed that all substrate adsorption is irreversible – that is, there is no desorption from the substrate. The corresponding mass of particles on the substrate, $\tilde {m}_s=\tilde {m}_s(\tilde r, \tilde t)$ , satisfies

(2.8) \begin{equation} \frac {\partial \tilde {m}_s}{\partial \tilde t} = \tilde {\rho }_{{s}}\tilde {k}_{{a}}\tilde {\phi }(\tilde r, 0, \tilde t), \end{equation}

where $\tilde {\rho }_{{s}}$ is the density of the colloid. Finally, it is assumed that the particles are initially uniformly distributed within the droplet, with no particles adsorbed onto the substrate, so that

(2.9) \begin{equation} \tilde \phi (\tilde r, \tilde z, 0) = \tilde \phi _{\textit{init}}, \quad \tilde m_s(\tilde r, 0) = 0. \end{equation}

To close the model, we require the evaporative flux $\tilde J$ . As stated above, we assume that the evaporation is primarily driven by a quasi-steady diffusion process, so that the molar vapour concentration $\tilde c=\tilde c(\tilde r, \tilde z, \tilde t)$ is determined by solving

(2.10) \begin{equation} {\nabla} ^2 \tilde c = 0 \quad \mbox{in the gas phase}, \end{equation}

subject to

(2.11) \begin{equation} \tilde c(\tilde r, \tilde h, \tilde t) = \tilde {c}_{\textit{sat}} \quad \mbox{for} \quad \tilde r\lt \tilde R, \quad \frac {\partial \tilde c}{\partial \tilde z}(\tilde r, 0, \tilde t) = 0 \quad \mbox{for} \quad \tilde r \gt \tilde R, \end{equation}

and such that

(2.12) \begin{equation} \tilde c \rightarrow \tilde {c}_\infty \quad \mbox{as} \quad \tilde r^{2}+\tilde z^{2}\rightarrow \infty , \end{equation}

where $\tilde {c}_{\textit{sat}}$ and $\tilde {c}_\infty$ are the constant saturation and ambient vapour concentrations, respectively. The required evaporative flux is then given by

(2.13) \begin{equation} \tilde J(\tilde r, \tilde t) = -\tilde {M}_\ell \tilde D_{v}\boldsymbol{\nabla }\tilde c\boldsymbol{\cdot }\boldsymbol{n}\big |_{\tilde h}, \end{equation}

where $\tilde M_\ell$ is the molar mass of the liquid vapour and $\tilde D_{v}$ is the constant diffusion coefficient of the vapour phase.

2.1. Non-dimensionalisation

Throughout this analysis, we concentrate on droplets that are thin, which may be expressed herein as

(2.14) \begin{equation} 0\lt \delta =\frac { \tilde V_{\textit{init}}}{ {\tilde R}^3 }\ll 1. \end{equation}

Under the thin assumption, we may linearise the mixed-boundary problem for the vapour concentration (2.10)–(2.12) onto $\tilde z = 0$ , retrieving the classical problem for the electrostatic potential outside a charged disk of radius $\tilde R$ , whose solution has been long known (see e.g. Weber (Reference Weber1873) and Copson (Reference Copson1947) for details): the relevant information for the present analysis is the evaporative flux, which is determined by the appropriate linearised version of (2.13) to be

(2.15) \begin{equation} \tilde J= \tilde {M}_\ell \tilde D_{v}\big(\tilde {c}_{\textit{sat}}-\tilde {c}_\infty \big)\frac {2}{\pi }\frac {1}{\sqrt {\tilde {R}^2-\tilde {r}^2}}. \end{equation}

Since we assume that the only way fluid is lost from the system is through evaporation, the liquid volume $\tilde V(\tilde t)$ satisfies

(2.16) \begin{equation} \frac {\mathrm{d} \tilde V}{\mathrm{d} \tilde t}= -\frac {4\tilde M_\ell \tilde D_{v}\big(\tilde {c}_{\textit{sat}}-\tilde {c}_\infty \big) \tilde R}{\tilde \rho } \quad \implies \quad \tilde V=\tilde V_{\textit{init}}-\frac {4\tilde M_\ell \tilde D_{v}\big(\tilde {c}_{\textit{sat}}-\tilde {c}_\infty \big) \tilde R \tilde t}{\tilde \rho }, \end{equation}

where the right-hand side of the differential equation is the linearised net evaporation rate, determined by integrating (2.15) over the droplet. Hence the lifetime of the droplet $\tilde t_{\!f}$ is

(2.17) \begin{equation} \tilde {t}_{\!f}=\frac {\tilde \rho \tilde V_{\textit{init}}}{4\tilde R\,\tilde M_\ell \tilde D_{v}\big(\tilde {c}_{\textit{sat}}-\tilde {c}_\infty \big)}. \end{equation}

Following previous studies (e.g. Moore et al. Reference Moore, Vella and Oliver2021; D’Ambrosio et al. Reference D’Ambrosio, Wray and Wilson2025b ), we non-dimensionalise (2.1)–(2.9), (2.15)–(2.17) by introducing the scalings

(2.18) \begin{align} \tilde t=\tilde {t}_{\!f} t, \quad (\tilde r, \tilde z,\tilde h)& = \tilde R( r, \delta z, \delta h), \quad (\tilde {u},\tilde w)=\tilde {U}(u,\delta w), \quad \tilde p =\tilde {p}_{{\textit{atm}}}+\frac {\tilde \mu \tilde {U}}{\delta ^2\tilde R}p, \nonumber\\ \tilde J & = \delta \rho \tilde {U}\!{J}, \quad \tilde \phi =\tilde {\phi }_{\textit{init}}\phi , \quad \tilde {m}_s = \delta \tilde R \tilde {\phi }_{\textit{init}}\tilde {\rho }_{{s}}m_s, \end{align}

where the velocity scale $\tilde {U} = {\tilde R}^2 \tilde M_\ell \tilde D_{v}(\tilde {c}_{\textit{sat}}-\tilde {c}_\infty )/(\tilde \rho \tilde V_{\textit{init}})$ and time scale have been selected as those driven by evaporation, and the pressure scale is chosen to balance the viscous terms in the radial momentum equation.

Under these scalings and the thin droplet assumption, it is straightforward to show in the usual manner (see e.g. Hocking (Reference Hocking1983), Deegan et al. (Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000) and Oliver et al. (Reference Oliver, Whiteley, Saxton, Vella, Zubkov and King2015), for details) that the gravity-free lubrication equations pertain:

(2.19) \begin{equation} \bar {u} = -\frac {h^2}{3}\frac {\partial p}{\partial r}, \quad p = -\frac {1}{{\textit{Ca}}}\frac {1}{r}\frac {\partial }{\partial r}\left (\!r\frac {\partial h}{\partial r}\!\right )\!, \quad\! 4\frac {\partial h}{\partial t}+\frac {1}{r}\frac {\partial }{\partial r}\left (\textit{rh} \bar {u} \right )+J= 0 \quad \mbox{for} \quad 0\lt r,\;t\lt 1, \end{equation}

where the depth-averaged radial velocity $\bar {u}=\bar {u}(r,t)$ is given by

(2.20) \begin{equation} \bar {u} = \frac {1}{h}\int _0^h u{\,\mathrm{d}} z \end{equation}

and the capillary number ${\textit{Ca}}$ is

(2.21) \begin{equation} {\textit{Ca}} = \frac {\tilde \mu \tilde U}{\delta ^3\tilde \sigma }. \end{equation}

As discussed in detail in § 2.2 and in previous studies (e.g. Moore et al. Reference Moore, Vella and Oliver2021), for a wide variety of real-world applications, the pertinent limit is that in which ${\textit{Ca}}\ll 1$ . Hence, to leading order in an asymptotic expansion as ${\textit{Ca}}\rightarrow 0$ , we find that

(2.22) \begin{equation} h(r,t) = \frac {\theta (t)}{2}(1-r^2), \quad \bar {u}(r,t) = \frac {4}{\pi \theta (t) r}\left (\frac {1}{\sqrt {1-r^2}}-(1-r^2)\right )\!, \end{equation}

where $\theta (t) = \theta _0(1-t)$ is the contact angle of the droplet in the thin-film limit, $\theta _0=\theta (0)=4/\pi$ is the initial contact angle of the droplet and the corresponding capillary pressure is $p(t) = 2\theta (t)/{\textit{Ca}}$ . Note that $\theta (1) = 0$ corresponds to the droplet lifetime. One notable property of (2.22) is that time is separable in both the free surface and depth-averaged velocity, the latter of which means that pathlines coincide with streamlines. Further, we see that time enters purely through the contact angle $\theta (t)$ : sometimes it is useful to exploit this fact and use $\theta$ as our effective time variable. We state in the following where we use this property.

With the flow and droplet shape to hand, the dimensionless advection–diffusion equation (2.5) becomes

(2.23) \begin{equation} 4\frac {\partial \phi }{\partial t}+u\frac {\partial \phi }{\partial r}+w\frac {\partial \phi }{\partial z}=\frac {1}{{\textit{Pe}}}\left [ \frac {1}{r}\frac {\partial }{\partial r}\left (r\frac {\partial \phi }{\partial r}\right ) +\frac {1}{\delta ^2}\frac {\partial ^2\phi }{\partial z^2}\right ] \quad \mbox{for} \quad 0\lt r,\ t\lt 1,\ 0\lt z\lt h, \end{equation}

where the Péclet number that represents the relative importance of advective flux and diffusive flux of particles is defined by

(2.24) \begin{equation} {\textit{Pe}}=\frac {\tilde R \tilde U}{\tilde D_{{s}}}. \end{equation}

In this study, we concentrate on the regime in which ${\textit{Pe}}\gtrsim 1$ , which is consistent with a wide range of different applications (see § 2.2 and Moore et al. (Reference Moore, Vella and Oliver2021)). In particular, this means that on the droplet scale, the colloidal diffusive flux is small compared with the advective flux, except in a small region near the pinned contact line, as we discuss shortly.

The dimensionless no-penetration condition on the free surface (2.6) is given by

(2.25) \begin{equation} \left (1+\delta ^2 \left (\dfrac {\partial h}{\partial r}\right )^2\right )^{-1/2}\left (\frac {\partial \phi }{\partial z}-\delta ^2 \frac {\partial h}{\partial r} \frac {\partial \phi }{\partial r}\right )=\delta ^2 {\textit{Pe}}\, \phi J \quad \mbox{on} \quad z = h, \end{equation}

while the substrate condition (2.7) becomes

(2.26) \begin{equation} \frac {\partial \phi }{\partial z}=\delta ^2 {\textit{Da}} \, \phi \quad \mbox{on} \quad z = 0, \end{equation}

where the importance of substrate adsorption is encoded in the Damköhler number defined by

(2.27) \begin{equation} {\textit{Da}}=\frac {\tilde R \tilde {k}_{{a}}}{\delta \tilde {D}_{{s}} }, \end{equation}

with ${\textit{Da}}\ll 1$ ( ${\textit{Da}}\gg 1$ ) indicating that substrate adsorption is negligible (dominant) compared with diffusive colloid transport. Finally, the initial condition is given by

(2.28) \begin{equation} \phi (r,z,0) = 1. \end{equation}

We seek an asymptotic expansion of the transport problem of the form

(2.29) \begin{equation} \phi =\phi _0+\delta ^2 \phi _1+\cdots \quad \mbox{as} \quad \delta \rightarrow 0. \end{equation}

Assuming that ${\textit{Pe}}, {\textit{Da}} \ll \delta ^{-2}$ , at leading order we find

(2.30) \begin{equation} \frac {\partial ^2\phi _0}{\partial z^2}=0 \quad \mbox{for} \quad 0\lt r,\; t\lt 1, \; 0\lt z\lt h, \end{equation}

subject to homogeneous Neumann boundary conditions on $z = 0, h(r,t)$ . Thus, to leading order, the volume fraction is independent of depth, so that $\phi _0=\phi _0(r,t)$ .

Treating both the Péclet and Damköhler numbers as formally $O(1)$ as $\delta \rightarrow 0$ , the $O(\delta ^2)$ -correction to the volume fraction satisfies

(2.31) \begin{equation} 4\frac {\partial \phi _0}{\partial t}+u\frac {\partial \phi _0}{\partial r}=\frac {1}{{\textit{Pe}}}\left [ \frac {1}{r}\frac {\partial }{\partial r}\left (r\frac {\partial \phi _0}{\partial r} \right )+\frac {\partial ^2\phi _1}{\partial z^2} \right ] \quad \mbox{for} \quad 0\lt r,\; t\lt 1, \; 0\lt z\lt h, \end{equation}

subject to

(2.32) \begin{equation} \frac {\partial \phi _1}{\partial z} = {\textit{Da}}\, \phi _0 \quad \mbox{on} \quad z = 0; \quad \frac {\partial \phi _1}{\partial z} = \frac {\partial h}{\partial r}\frac {\partial \phi _0}{\partial r}+{\textit{Pe}}\, \phi _0 J \quad \mbox{on} \quad z = h(r,t). \end{equation}

Whence, upon integrating (2.31) across the droplet and employing the lubrication equations (2.19), we deduce

(2.33) \begin{equation} \frac {\partial }{\partial t}(h\phi _0)+\frac {1}{4r}\frac {\partial }{\partial r}\left [r\left (h\phi _0\bar {u}-\frac {1}{{\textit{Pe}}}h\frac {\partial \phi _0}{\partial r}\right )\right ]+\phi _0\frac {{\textit{Da}}}{4{\textit{Pe}}}=0 \end{equation}

for $0\lt r,\ t\lt 1$ . This must be solved subject to the condition of initial uniform distribution of solute:

(2.34) \begin{equation} \phi _0(r,0) = 1, \end{equation}

as well as relevant boundary conditions in the radial direction. The pertinent choice on physical grounds is to specify no-flux boundary conditions through the axis of symmetry and the contact line, so that

(2.35) \begin{equation} r\left (h\phi _0\bar {u}-\frac {1}{{\textit{Pe}}}h\frac {\partial \phi _0}{\partial r}\right ) = 0 \quad \mbox{at} \quad r = 0,1. \end{equation}

After non-dimensionalisation, the corresponding growth in the mass of particles on the substrate is given by $m_s = m_{s,0} + o(1)$ as $\delta \rightarrow 0$ , where

(2.36) \begin{equation} \frac {\partial m_{s,0}}{\partial t} = \frac {{\textit{Da}}}{4{\textit{Pe}}} \phi _0, \quad m_{s,0}(r,0) = 0. \end{equation}

Finally, integrating (2.33) and (2.36) over the footprint of the droplet, applying the conditions (2.34)–(2.35) and summing yields global conservation of the particle mass:

(2.37) \begin{equation} 1 = 2\pi \int _{0}^{1} \textit{rh}\phi _0 \,{\mathrm{d}} r + 2\pi \int _{0}^{1} {\textit{rm}}_{s,0}\,{\mathrm{d}} r, \end{equation}

where the term on the left-hand side gives the total initial particle mass and the integrals on the right-hand side give, respectively, the total mass of particles remaining in suspension and adsorbed onto the substrate.

Henceforth, we drop the ‘ $0$ ’ subscript notation denoting the leading-order variables for brevity.

2.2. Parameter estimation

The remainder of our analysis focuses on solutions to (2.33)–(2.37). Naturally, this encompasses a wide parameter space spanned by the Péclet and Damköhler numbers. To that end, we first discuss realistic ranges for these numbers and focus our attention on these regimes. We also revisit some previous assumptions made herein, namely that  $\delta , \; {\textit{Ca}}, \; {{Bo}}\ll 1$ .

A limiting factor in this endeavour is available data for the substrate adsorption velocity  $\tilde {k}_{{a}}$ . Adsorption kinetics is a complex topic in its own right and may depend on factors such as electrostatic interactions; particle geometry and properties; and intermolecular forces. As a result, reliable experimental data can be difficult to obtain. In this section, we discuss some of the pertinent literature in the context of the present mathematical model.

First, we discuss how we may estimate the substrate adsorption velocity from experimental data in the literature. Perhaps the simplest theory available for estimating the adsorption rate of particles from fluid flow onto a ‘collector’ (i.e. a substrate) is the Smoluchowski–Levich approximation which assumes that, in the absence of colloidal and hydrodynamic interactions, deposition is controlled by convective–diffusive transport. The substrate adsorption velocity may then be expressed as

(2.38) \begin{equation} \tilde {k}_{{a}}=\dfrac {\tilde {D}_{{s}}}{\tilde {\delta }_D} \end{equation}

(see e.g. Adamczyk Reference Adamczyk2019), where $\tilde {\delta }_D$ is the thickness of the diffusive boundary layer close to the collector. In general, $\tilde {\delta }_D$ depends on the collector/flow configuration and is inversely proportional to the dimensionless mass transfer Sherwood number which has been estimated for a variety of ideal collectors (i.e. assuming the collector is a perfect sink) in different flow configurations (see e.g. Adamczyk et al. Reference Adamczyk, Dabros, Czarnecki and Van De Ven1983; Elimelech Reference Elimelech1994). For the present analysis of an evaporating droplet we assume that the collector/flow configuration is that of a spherical collector in uniform flow, and so the substrate adsorption velocity may then be estimated by

(2.39) \begin{equation} \tilde {k}_{{a}}\approx \dfrac {(\delta \tilde {U})^{1/3}\tilde {D}_{{s}}^{2/3}}{\tilde {R}^{2/3}} \end{equation}

(see e.g. Adamczyk et al. Reference Adamczyk, Dabros, Czarnecki and Van De Ven1983; Adamczyk Reference Adamczyk2002), where $\tilde {U}$ is the characteristic radial velocity within the droplet from (2.18) and $\delta \tilde {U}$ is therefore the approach velocity of a particle towards the substrate.

However, as previously mentioned, the Smoluchowski–Levich approximation neglects several factors, including particle–substrate repulsion forces. When there exists an energy barrier between the particle and the substrate an alternative theory may be used to estimate $\tilde {k}_{{a}}$ , namely interaction-force boundary-layer theory (see e.g. Spielman & Friedlander Reference Spielman and Friedlander1974; Adamczyk Reference Adamczyk2002; Adamczyk et al. Reference Adamczyk, Siwek, Weroński and Jaszczółt2003), sometimes also called surface boundary-layer theory. Specifically, assuming that substrate adsorption is a linear and irreversible process, the constant rate of adsorption between a spherical particle and a planar substrate may be approximated by the expression

(2.40) \begin{equation} \tilde {k}_{{a}}=\tilde {D}_{{s}}\left (\int _{\tilde \delta _{{m}}}^{\infty }{\left (e^{\tilde {V}_{{ t}}(\tilde d)/(\tilde k_{{B}}\tilde T)}-1\right )f(\tilde d)\,\text{d}\tilde d}\right )^{-1}\!, \end{equation}

where $\tilde V_{{t}}(\tilde d)$ , $\tilde k_{{B}}$ , $\tilde T$ , $f(\tilde d)$ , $\tilde d$ and $\tilde \delta _{{m}}$ denote the total interaction energy potential between the particle and the substrate, the Boltzmann constant, temperature, the hydrodynamic correction function, the distance between the particle and the substrate and the location of the primary minimum of the net interaction energy potential, respectively. In particular, the hydrodynamic correction function may be approximated by $f(\tilde d)=1+\tilde r_{{p}}/\tilde d$ (Adamczyk Reference Adamczyk2002), where $\tilde r_{{p}}$ is the radius of the particle, and which describes the resistance that arises due to the proximity of a particle to the substrate and therefore acts to reduce the particle diffusion coefficient relative to the bulk diffusivity of particles in the fluid, $\tilde {D}_{{s}}$ . The net interaction energy potential $\tilde V_{{t}}$ between a particle and the substrate may depend on a variety of long- and short-range interactions that balance attractive and repulsive forces. Here we follow Zigelman & Manor (Reference Zigelman and Manor2018a ) and Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) by adopting Derjaguin–Landau–Verwey–Overbeek (DLVO) theory in which we assume that the dominant forces at play are electrostatic and van der Waals (attractive) forces. The net interaction energy potential may then be expressed as $\tilde V_{{t}}=\tilde V_{{el}}+\tilde V_{{v{\rm d}W}}$ with

(2.41) \begin{align} \tilde V_{{el}}=64\pi \tilde \varepsilon _0\tilde \varepsilon \left (\dfrac {\tilde k_{{B}} \tilde T}{\tilde e}\right )^2\tanh {\left (\dfrac {\tilde {\mathrm{e}} \tilde \psi _{{p}}}{4\tilde k_{{B}} \tilde T}\right )}\tanh {\left (\dfrac {\tilde {\mathrm{e}}\tilde \psi _{{s}}}{4\tilde k_{{B}}\tilde T}\right )}e^{-\tilde \kappa \tilde d} \quad \text{and} \quad \tilde V_{{v{\rm d}W}}=-\dfrac {\tilde A\tilde r_{{p}}}{6 \tilde d}\\[-20pt]\nonumber \end{align}

(see e.g. Zigelman & Manor Reference Zigelman and Manor2018a ; Bridonneau et al. Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020), where $\tilde \varepsilon _0\tilde \varepsilon$ , $\tilde \kappa$ , $\tilde e$ , $\tilde \psi _p$ , $\tilde \psi _s$ and $\tilde A$ are the total permittivity of the liquid, the reciprocal of the Debye length, the elementary charge, the surface potential of the particle, the surface potential of the substrate and the Hamaker constant, respectively. With the Smoluchowski–Levich approximation and the interaction-force boundary-layer theory descriptions in hand, we are ready to estimate the substrate adsorption velocity from experimental data in the literature.

First, we consider the experimental data of Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) in which the deposition of silica nanoparticles from water droplets evaporating on gold-coated silicon wafer substrates is studied. In particular, the initial particle concentration is varied across four orders of magnitude, and the charges of the particles and the substrate are altered through different coatings and/or treatments prior to evaporation to study the effect of particle–substrate interaction on the deposit. Specifically, they found that the deposit shape remained the same regardless of the particle–substrate interaction in the studied range of DLVO forces, indicating that the aforementioned Smoluchowski–Levich approximation is appropriate to determine the substrate adsorption velocity for this study.

We look to determine our model parameters for two different experiments, specifically for droplets containing $1\,\text{g}\,\text{l}^{-1}$ of either negatively charged untreated silica nanoparticles or positively charged aluminium-coated silica nanoparticles evaporating on an oppositely charged gold substrate, respectively. These cases have been chosen since, whilst Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) found that the qualitative shape of the deposit remained the same regardless of the particle–substrate interaction, oppositely charged particles and substrates are the ideal configuration to use the Smoluchowski–Levich approximation since repulsion forces exist at small separations when the particles and the substrate have the same charge. For both experiments, the droplet has an initial volume $\tilde {V}_{\textit{init}}\approx 0.5$ μl. The initial contact angles of the droplets are ${\approx} 60.1^\circ$ and $54.4^\circ$ , from which we may estimate the initial contact radii to be $\tilde {R}\approx 0.79$ and $0.83\,\text{mm}$ , for droplets containing untreated and aluminium-coated silica nanoparticles, respectively. Hence, the aspect ratios of the droplets are $\delta \approx 1.01$ and $0.88$ . While these values of $\delta$ push the realms of the thin-film assumption initially, due to the contact line pinning, the thin droplet assumption becomes stronger as the droplets evaporate and the contact angles decrease. Furthermore, Larsson & Kumar (Reference Larsson and Kumar2022) demonstrate that the lubrication approximation is often a reasonable assumption in regimes where it strictly may not be asymptotically justified. The characteristic radial velocities are estimated to be $\tilde {U}\approx 4.14\times 10^{-7}$ and $4.55\times 10^{-7}\,\text{m}\,\text{s}^{-1}$ . The capillary and Bond numbers are therefore ${\textit{Ca}}\approx 5.12\times 10^{-9}$ and $5.62\times 10^{-9}$ and ${{Bo}}\approx 0.085$ and $0.093$ , respectively, which justifies our assumptions that surface tension dominates viscosity in the liquid flow and that gravity is negligible. The untreated and aluminium-coated silica nanoparticles have radii $\tilde {r}_{{p}}=37.5$ and $45\,\text{nm}$ and so the diffusivity of the particles in water may be determined from the Stokes–Einstein relation to be $\tilde {D}_{{s}}=6.55\times 10^{-12}$ and $5.46\times 10^{-12}\,\text{m}^2\,\text{s}^{-1}$ , respectively, so that the Péclet numbers are then ${\textit{Pe}}\approx 50.0$ and $69.1$ . Finally, the Smoluchowski–Levich approximation described above for the substrate adsorption velocity yields $\tilde {k}_{{a}}\approx 3.06\times 10^{-8}$ and $2.59\times 10^{-8}\,\text{m}\,\text{s}^{-1}$ , and so the Damköhler numbers are ${\textit{Da}}\approx 3.66$ and $4.47$ .

Next, we consider experimental data of Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) in which the deposition of titania particles from water droplets of different pH values evaporating on glass substrates is studied. In particular, as outlined by Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010), both the electrostatic and van der Waals forces (or the DLVO force) between a particle and the substrate vary with the acidity of the base fluid, and a switch from particle–substrate attraction to repulsion is observed as the pH value of the fluid increases.

We look to determine our model parameters from values provided for two different experiments, namely when $\text{pH}=2.8$ and $\text{pH}=11.7$ . Specifically, when $\text{pH}=2.8$ the particles and the substrate are oppositely charged and the DLVO force is attractive, and when $\text{pH}=11.7$ the particles and the substrate are both negatively charged and the DLVO force is repulsive. The droplets have initial volume $\tilde {V}_{\textit{init}}\approx 5\,\text{nl}$ . The initial contact radii of the droplets are $\tilde {R}\approx 345$ and $410$ μm, from which we may estimate the initial contact angles to be ${\approx} 8.85^\circ$ and $5.29^\circ$ , for the attractive and repulsive cases, respectively. Hence, the aspect ratios of the droplets are $\delta \approx 0.12$ and $0.07$ , which are within our $0\lt \delta \ll 1$ assumption. The characteristic radial velocity scales are estimated to be $\tilde {U} \approx 7.22\times 10^{-6}$ and $1.02\times 10^{-5}\,\text{m}\,\text{s}^{-1}$ . The capillary and Bond numbers for this example are therefore ${\textit{Ca}}\approx 8.93\times 10^{-8}$ and $1.26\times 10^{-7}$ and ${{Bo}}\approx 0.016$ and $0.023$ , respectively. The titania particles have radii $\tilde {r}_{{p}}=12.5\,\text{nm}$ and so the diffusivity of the particles in water may be determined from the Stokes–Einstein relation to be $\tilde {D}_{{s}} \approx 1.96\times 10^{-11}\,\text{m}\,\text{s}^{-1}$ , so that the Péclet numbers are ${\textit{Pe}}\approx 127$ and $213$ . Finally, for the experiment in which $\text{pH}=2.8$ , the DLVO force is attractive and we may use the Smoluchowski–Levich approximation to estimate the substrate adsorption velocity to be $\tilde {k}_{{a}}\approx 1.42\times 10^{-7}\,\text{m}\,\text{s}^{-1}$ , which gives ${\textit{Da}}\approx 20.5$ . For the experiment in which $\text{pH}=11.7$ , the DLVO force is repulsive and we may use interaction-force boundary-layer theory to estimate $\tilde {k}_{{a}}\approx 2.86\times 10^{-8}\,\text{m}\,\text{s}^{-1}$ and hence ${\textit{Da}}\approx 8.23$ .

We now turn to examples in the literature for which the substrate adsorption velocity is explicitly estimated. The present model of substrate adsorption as represented by (2.7) is based on pioneering modelling and experiments on the kinetics of bovine serum albumin (BSA) adsorption and the values for $\tilde {k}_{{a}}$ determined therein are used in the examples studied experimentally and numerically in Siregar (Reference Siregar2012) and Siregar et al. (Reference Siregar, Kuerten and Van der Geld2013). We take the value of $\tilde {k}_{{a}} \approx 2\times 10^{-8}$ m s $^{-1}$ for BSA in a HEPES buffer from Kurrat, Prenosil & Ramsden (Reference Kurrat, Prenosil and Ramsden1997) for a reference value of the substrate adsorption velocity. The diffusivity of BSA is given in Kurrat et al. (Reference Kurrat, Prenosil and Ramsden1997) to be $\tilde {D}_{{s}} \approx 3.26\times 10^{-11}$ m $^2$ s $^{-1}$ .

The first example in Siregar et al. (Reference Siregar, Kuerten and Van der Geld2013) considers a droplet based on data from Hu & Larson (Reference Hu and Larson2002), for an evaporating water droplet with an initial volume $\tilde {V}_{\textit{init}} \approx 0.60$ μl and an initial contact radius $\tilde {R} \approx 1$ mm. The aspect ratio parameter is thus $\delta \approx 0.597$ . The evaporation velocity scale is estimated to be $\tilde {U}\approx 2.19\times 10^{-7}$ m s $^{-1}$ . Hence, the capillary and Bond numbers for this example are ${\textit{Ca}}\approx 1.25\times 10^{-8}$ and ${{Bo}}\approx 0.13$ , respectively, which again justifies our assumptions that surface tension dominates viscosity in the liquid flow and that gravity is negligible. The Péclet number is given by ${\textit{Pe}}\approx 6.72$ , while the Damköhler number is ${\textit{Da}} \approx 1.03$ .

The second example in Siregar et al. (Reference Siregar, Kuerten and Van der Geld2013) is based on evaporation data from Golovko, Butt & Bonaccurso (Reference Golovko, Butt and Bonaccurso2009) and considers a $\tilde {V}_{\textit{init}} = 66.2$ pl water droplet of initial contact radius $\tilde {R} = 37.3$ μm. The evaporation velocity scale is estimated to be $\tilde {U}\approx 7.04\times{} 10^{-6}$ m s $^{-1}$ . The capillary and Bond numbers are therefore ${\textit{Ca}}\approx 4.1\times 10^{-8}$ and ${{Bo}} \approx 1.89\times 10^{-4}$ , justifying the supposition that surface tension dominates the droplet dynamics. On the other hand, the aspect ratio $\delta \approx 1.28$ for this example, which is due to the large initial contact angle ${\approx} 70^\circ$ . The above caveats about the role of $\delta$ pertain once again. The Péclet number is ${\textit{Pe}} \approx 8.05$ , while the Damköhler number is ${\textit{Da}}\approx 1.79\times 10^{-2}$ .

As a final example, we turn to a different colloid, namely polystyrene latex particles, whose adsorption onto mica is studied experimentally in Adamczyk et al. (Reference Adamczyk, Siwek, Weroński and Jaszczółt2003). The authors compare heterogeneous and homogeneous surfaces, with adsorption velocities in the range $\tilde {k}_{{a}} \approx (8.9{-}9.7)\times 10^{-8}$ m s $^{-1}$ , while the diffusion coefficient is estimated from figure 2(b) and equation (2) therein to be $\tilde {D}_{{s}} \approx 1.02\times 10^{-12}$ m $^2$ s $^{-1}$ . For reference, we consider the water droplet from the first example based on the data in Hu & Larson (Reference Hu and Larson2002) containing such particles, leading to ${\textit{Pe}} \approx 219$ and ${\textit{Da}} \approx 154$ .

From the above examples, it is of interest to consider a range of possible Péclet and Damköhler numbers. However, given that we are primarily concerned with the influence of substrate adsorption on coffee-ring formation, we focus our asymptotic analysis on the regime for which ${\textit{Pe}} \gg 1$ , since it is well known that it is the dominance of advection in the droplet bulk and its competition with diffusion near the contact line that leads to the nascent coffee ring (Moore et al. Reference Moore, Vella and Oliver2021) (see § 3). While this assumption is consistent with the majority of the studies referenced previously, there are examples for which the Péclet number approaches order unity, so we return to such cases when considering our numerical results (see § 4).

In terms of the Damköhler number, we note that while this can vary from large to small in the examples above, in each case we have ${\textit{Da}}\lesssim {\textit{Pe}}$ . To this end, we may introduce the parameter $\mathcal{V}$ defined by

(2.42) \begin{equation} \frac {2\mathcal{V}}{\pi } := \frac {{\textit{Da}}}{{\textit{Pe}}} = \frac {\tilde {k}_{{a}}}{\delta \tilde {U}}, \end{equation}

where the coefficient $2/\pi$ has been introduced for notational convenience in the following. This is the ratio between the substrate adsorption velocity $\tilde {k}_{{a}}$ and the vertical velocity across the droplet. Hence, if $\mathcal{V}\lesssim 1$ , we are still in a regime where evaporation is a dominant effect, while if $\mathcal{V}\gg 1$ , substrate adsorption dominates evaporation. Since the latter regime is beyond the scope of the present paper, where our primary interest is in ring formation, we focus on the regime for which ${\textit{Da}}\lesssim {\textit{Pe}}$ henceforth.

In what follows, we perform a matched asymptotic analysis of (2.33)–(2.37) in the advection-dominated regime in § 3 to both gain physical insight and act as a validation tool for our numerical simulations; the numerical methodology is outlined in Appendix A. In the interests of completeness, we briefly discuss the regime in which advection is weaker in § 4.

3. Advection-dominated transport, ${\textit{Pe}}\boldsymbol{\gg 1}$

As discussed above, we concentrate on advection-dominated transport herein and, for the purposes of our analysis, we assume that $\mathcal{V} = O(1)$ as ${\textit{Pe}}\rightarrow \infty$ . Regimes with weaker substrate adsorption will be recovered as regular sublimits of the analysis, as we shall see presently.

The asymptotic structure follows that of Moore et al. (Reference Moore, Vella and Oliver2021), whereby in the bulk of the droplet, we may neglect colloidal diffusion in (2.33), which is only a leading-order effect in the vicinity of the contact line, where the volume fraction increases sharply due to advection.

3.1. Outer problem

In the droplet bulk, the two physical effects influencing the volume fraction distribution are advection due to evaporation and settling due to substrate adsorption. To leading order, the volume fraction satisfies the first-order linear equation

(3.1) \begin{equation} \frac {\partial }{\partial t}(h\phi )+\frac {1}{4r}\frac {\partial }{\partial r}\left (r h\phi \bar {u}\right )+\phi \frac {\mathcal{V}}{2\pi }=0 \quad \mbox{for} \quad 0\lt r,\; t\lt 1, \end{equation}

which can be simplified using the lubrication equation and by switching our time variable to the more convenient contact angle using the transformation $\theta = \theta _0(1-t)$ , finding

(3.2) \begin{equation} \frac {\partial \phi }{\partial \theta } - \frac {\bar {u}}{4\theta _0}\frac {\partial \phi }{\partial r} = -\frac {\phi }{2\pi \theta _0 h}\left (\frac {1}{\sqrt {1-r^2}}-\mathcal{V}\right ) \quad \mbox{for} \quad 0\lt r\lt 1,\; 0\lt \theta \lt \theta _0. \end{equation}

This must be solved subject to the initial condition (2.34). This problem is exactly that studied in D’Ambrosio et al. (Reference D’Ambrosio, Wray and Wilson2025b ) and may be solved using the method of characteristics, yielding

(3.3) \begin{equation} \phi =\left (\frac {\theta }{\theta _0}\right )^{-1/4-\mathcal{V}/8}\frac {\xi }{\xi _0}\left (\frac {1-\xi }{1-\xi _0}\right )^{\mathcal{V}/2}\mbox{exp}\left [\frac {\mathcal{V}}{\sqrt {3}}\mbox{arctan}\left (\frac {\sqrt {3}(\xi _0-\xi )}{2\xi \xi _0+\xi +\xi _0+2}\right )\right ]\!, \end{equation}

where

(3.4) \begin{equation} \xi (r,\theta ) = \left (1 - \left (\frac {\theta }{\theta _0}\right )^{3/4}\big(1-\xi _0^3\big)\right )^{1/3}, \quad \xi _0 (r)= \sqrt {1-r^2}. \end{equation}

In particular, we note that

(3.5) \begin{equation} \phi (r,\theta )\sim \dfrac {\gamma (\theta )}{\sqrt {1-r}} \quad \text{as} \quad 1-r\to 0^{+} , \end{equation}

where

(3.6) \begin{equation} \gamma (\theta ) = \left (\!\frac {\theta }{\theta _0}\!\right )^{-1/4-\mathcal{V}/8}\frac {\zeta (\theta )(1-\zeta (\theta ))^{\mathcal{V}/2}}{\sqrt {2}}\mbox{exp}\!\left (\!-\frac {\mathcal{V}}{\sqrt {3}}\mbox{arctan}\!\left (\!\frac {\sqrt {3}\zeta (\theta )}{\zeta (\theta )+2}\!\right )\!\right )\!, \!\quad\! \zeta (\theta ) = \xi (1,\theta ). \end{equation}

Hence, the leading-order outer volume fraction becomes unbounded local to the contact line, necessitating a reconsideration of particle diffusion in a boundary layer.

We can use the expression for the behaviour of the volume fraction as $1-r\rightarrow 0^+$ , from (3.5) and (3.6), to calculate the accumulated mass flux into the contact line, $\mathcal{M}(\theta )$ , as given by

(3.7) \begin{equation} \mathcal{M}(\theta ) = -\frac {1}{4\theta _0}\int _{\theta _0}^{\theta } \left .(rh\bar {u}\phi )\right |_{r=1^{-}}{\,\mathrm{d}} \bar {\psi } = \frac {\chi }{2\theta _0}\int _{\theta }^{\theta _0} \gamma (\psi )\,{\mathrm{d}}\psi , \end{equation}

where $\chi = \sqrt {2}/\pi$ is the coefficient of the inverse square-root singularity in the diffusive evaporative flux, which arises from the local velocity expansion $\bar {u} \sim 2\chi /\theta \sqrt {1-r}$ as $1-r\rightarrow 0^+$ . We have introduced this notation to be consistent with previous analyses of the nascent coffee ring (Moore, Vella & Oliver Reference Moore, Vella and Oliver2022; Moore & Wray Reference Moore and Wray2023), in which the equivalent accumulated mass flux into the contact line in the absence of substrate adsorption is

(3.8) \begin{equation} \mathcal{M}_{\textit{MVO}}(\theta ) = \frac {\sqrt {2}\chi }{4}\left (1-\left (\frac {\theta }{\theta _{0}}\right )^{3/4}\right )^{4/3}\!. \end{equation}

We compare (3.7) and (3.8) in figure 2, where we see that the presence of substrate adsorption decreases $\mathcal{M}(\theta )$ relative to $\mathcal{M}_{\textit{MVO}}(\theta )$ , with mass now lost to the system via deposition. As the relative importance of substrate adsorption increases, this effect is amplified. We also show on the graph the small-time asymptotic form of $\mathcal{M}(\theta )$ for $\mathcal{V}\gt 0$ , namely

(3.9) \begin{align} \mathcal{M}(\theta ) \sim \frac {3^{4/3}\chi }{2^{25/6}\theta _0^{4/3}}(\theta _0-\theta )^{4/3}\left [1 - \frac {3^{1/3}2^{4/3}\mathcal{V}}{5\theta _0^{1/3}}(\theta _0-\theta )^{1/3} + \frac {\mathcal{V}^2}{3^{1/3}2^{4/3}\theta _0^{2/3}}(\theta _0-\theta )^{2/3}\right ]\\[-20pt]\nonumber \end{align}

as $\theta _0-\theta \rightarrow 0^+$ and the late-time expansion

(3.10) \begin{equation} \mathcal{M}(\theta ) \sim \frac {\chi }{2\theta _0}\left [\int _{0}^{\theta _0}\gamma (\psi )\,{\mathrm{d}} \psi - \frac {2\sqrt {2}e^{-\mathcal{V}\pi /6\sqrt {3}}}{3^{\mathcal{V}/2}(\mathcal{V}+6 )}\left (\frac {\theta }{\theta _0}\right )^{(\mathcal{V}+6)/8}\right ]\quad \mbox{as} \quad \theta \rightarrow 0^+. \end{equation}

Figure 2. Accumulated mass flux into the contact line $\mathcal{M}(\theta )$ for $\mathcal{V} = 0.25, 1, 4, 8$ (greyscales) and the equivalent coefficient in the no-substrate adsorption problem $\mathcal{M}_{\textit{MVO}}(\theta )$ (blue). The dashed red curves represent the small-time and large-time limits (3.9) and (3.10).

The reduction in the flux of particles to the contact line due to the mass adsorbed onto the substrate is given by integrating (2.36) to leading order in the bulk, yielding

(3.11) \begin{equation} m_s(r,\theta ) = \frac {\mathcal{V}}{2\pi \theta _0}\int _{\theta }^{\theta _0} \phi (r,\psi ) \,{\mathrm{d}} \psi . \end{equation}

Notably, as we approach the contact line, this is unbounded, with

(3.12) \begin{equation} m_s(r,\theta ) \sim \left (\frac {\mathcal{V}}{2\pi \theta _0}\int _{\theta }^{\theta _0} \gamma (\psi ) \,{\mathrm{d}} \psi \right )\frac {1}{\sqrt {1-r}} = \frac {\mathcal{V}\mathcal{M}(\theta )}{\pi \chi \sqrt {1-r}} \quad \mbox{as} \quad 1-r\rightarrow 0^+. \end{equation}

With reference to figure 2, it is worth stressing that the transition from $\mathcal{V} = 0$ to $\mathcal{V}\gt 0$ appears to be regular in terms of the accumulated mass flux: indeed, an asymptotic expansion of the outer solution in the limit $\mathcal{V} \rightarrow 0$ retrieves the equivalent outer solution without substrate adsorption (see e.g. Deegan et al. Reference Deegan, Bakajin, Dupont, Huber, Nagel and Witten2000; Moore et al. Reference Moore, Vella and Oliver2021). Formally, perturbations due to substrate adsorption are dominant over colloidal diffusion in the bulk provided that $\mathcal{V} \gg {\textit{Pe}}^{-1}$ as ${\textit{Pe}}\rightarrow \infty$ . If $\mathcal{V}\ll {\textit{Pe}}^{-1}$ , perturbations due to diffusion dominate those due to substrate adsorption in the droplet bulk, which is beyond the scope of the present analysis.

3.2. Inner region

We now turn to the boundary layer near the contact line. Note that we continue to use the contact angle as our time variable in this section. Following Moore et al. (Reference Moore, Vella and Oliver2021), we introduce local variables

(3.13) \begin{equation} r = 1 - \frac {R}{{\textit{Pe}}^2}, \quad \bar {u} = {\textit{Pe}} \bar {U}, \quad h = \frac {H}{{\textit{Pe}}^2}, \quad \phi = {\textit{Pe}}^b \varPhi , \quad m_s = {\textit{Pe}}^b M_s, \end{equation}

where $b\gt 0$ is to be determined from conservation of mass considerations shortly. Then, to leading order as ${\textit{Pe}}\rightarrow \infty$ , the dominant terms in (2.33) are advection and diffusion, so that

(3.14) \begin{equation} \frac {\partial }{\partial R} \left (\bar {U}\!H\kern-0.5pt\varPhi + H\frac {\partial \varPhi }{\partial R}\right ) = 0 \quad \mbox{for} \quad R\gt 0, \; 0\lt \theta \lt \theta _0, \end{equation}

subject to the leading-order no-flux condition at the contact line,

(3.15) \begin{equation} \bar {U}\!H\kern-0.5pt\varPhi + H\frac {\partial \varPhi }{\partial R} = 0 \quad \mbox{on} \quad R=0, \, 0\lt \theta \lt \theta _0, \end{equation}

together with the leading-order forms of the free surface and velocity as found from (2.22).

The leading-order boundary-layer solution is thus

(3.16) \begin{equation} \varPhi (R,\theta ) = A(\theta )\mbox{exp}\left (-\frac {4\chi }{\theta }\sqrt {R}\right )\!, \end{equation}

where $A(\theta )$ must be determined by matching. The exponential form of the solution is exactly the same as that presented in Moore et al. (Reference Moore, Vella and Oliver2022) in the no-substrate adsorption regime: the difference between the two solutions will solely come from the coefficient $A(\theta )$ , as we discuss further shortly.

Having determined the leading-order inner volume fraction, the leading-order adsorbed mass is then given by

(3.17) \begin{equation} M_s(R,\theta ) = \frac {\mathcal{V}}{2\pi \theta _0}\int _{\theta }^{\theta _0} A(\psi )\mbox{exp}\left (-\frac {4\chi }{\psi }\sqrt {R}\right )\,{\mathrm{d}} \psi . \end{equation}

3.2.1. Matching using the conservation of particle mass

In order to match between the leading-order outer and inner solutions, we must now turn to conservation of particle mass (2.37). Each of the integrals on the right-hand side of (2.37) may be split into contributions from the outer and inner solutions by introducing a parameter $0\lt {\textit{Pe}}^{-2}\ll \varepsilon \ll 1$ . For example, the suspension mass may be written as

(3.18) \begin{align} \int _{0}^{1} (rh\phi ) (r,\theta )\,{\mathrm{d}} r = \int _{0}^{1-\varepsilon } (rh\phi )_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + {\textit{Pe}}^{b-4} \int _{0}^{\varepsilon {\textit{Pe}}^2} \!\!\left (\!1-\frac {R}{{\textit{Pe}}^2}\!\right )\! (H\kern-0.5pt\varPhi)(R,\theta )\,{\mathrm{d}} R,\\[-20pt]\nonumber \end{align}

where $\phi _{\textit{outer}}$ is given by (3.3). Then, taking the limit $\varepsilon {\textit{Pe}}^2\rightarrow \infty$ and ${\textit{Pe}}\rightarrow \infty$ , we find to leading order

(3.19) \begin{align} \int _{0}^{1} (rh\phi )(r,\theta )\,{\mathrm{d}} r = \int _{0}^{1} (rh\phi )_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + {\textit{Pe}}^{b-4} \int _{0}^{\infty } (H\kern-0.5pt\varPhi) (R,\theta )\,{\mathrm{d}} R.\\[-20pt]\nonumber \end{align}

A similar argument for the adsorbed mass yields

(3.20) \begin{equation} \int _{0}^{1} {\textit{rm}}_s(r,\theta )\,{\mathrm{d}} r = \int _{0}^{1} ({\textit{rm}}_s)_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + {\textit{Pe}}^{b-2} \int _{0}^{\infty } M_s(R,\theta )\,{\mathrm{d}} R, \end{equation}

where $m_{s,{outer}}$ is given by (3.11).

Upon substituting these into (2.37), we have

(3.21) \begin{align} \displaystyle \frac {1}{2\pi } = \displaystyle \int _{0}^{1} \!(rh\phi + {\textit{rm}}_s)_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + {\textit{Pe}}^{b-4} \!\int _{0}^{\infty } \!(\!H\kern-0.5pt\varPhi \!) (\!R,\theta )\,{\mathrm{d}} R + {\textit{Pe}}^{b-2} \!\int _{0}^{\infty }\! M_s(\!R,\theta )\,{\mathrm{d}} R.\\[-20pt]\nonumber \end{align}

Now, since some mass is advected to the contact line (cf. (3.7)), for conservation of mass to be satisfied to leading order as ${\textit{Pe}}\rightarrow \infty$ , we cannot only consider the outer region and must include one of the inner-region terms. Here, the correct dominant balance is to choose $b = 2$ , so that leading-order conservation of mass gives

(3.22) \begin{equation} \frac {1}{2\pi } = \int _{0}^{1} (rh\phi + {\textit{rm}}_s)_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + \int _{0}^{\infty } M_s(R,\theta )\,{\mathrm{d}} R. \end{equation}

This is different from the equivalent formulation without substrate adsorption as discussed in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022), where we must take $b = 4$ . Substrate adsorption thus acts to reduce the size of the volume fraction in the boundary layer in the present regime.

If we differentiate (3.22) with respect to $\theta$ , we find that

(3.23) \begin{equation} 0 = -\frac {1}{4}\int _{0}^{1} \frac {\partial }{\partial r} \left ((rh\phi \bar {u})_{\textit{outer}}\right ) \,{\mathrm{d}} r + \frac {\mathcal{V}}{2\pi } \int _{0}^{\infty } \varPhi (R,\theta )\,{\mathrm{d}} R, \end{equation}

where we have used the governing advection equation for the outer solution and the inner adsorption equation. Whence, utilising the no-flux condition at the droplet centre and the known solution for $\varPhi (R,\theta )$ given in (3.16), we deduce that

(3.24) \begin{equation} A(\theta ) = -\frac {16\pi }{\mathcal{V}}\frac {\theta _0\chi ^2}{\theta ^2}\frac {{\,\mathrm{d}}\mathcal{M}}{{\,\mathrm{d}}\theta } = \frac {8\pi \chi ^3\gamma (\theta )}{\mathcal{V}\theta ^2}. \end{equation}

The expression (3.24) has been presented in a slightly unusual form to highlight where physical quantities in the system such as $\mathcal{M}(\theta )$ , $\chi$ and $\theta$ naturally occur. This is for ease of interpreting the results and for comparison with other regimes where substrate adsorption may be more or less important, or indeed, comparisons with more general geometries.

We display the coefficient $A(\theta )$ in figure 3, where we also include for reference the result for the zero-substrate adsorption regime in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022), which is

(3.25) \begin{equation} A_{\textit{MVO}}(\theta ) = \frac {64\chi ^4\mathcal{M}_{\textit{MVO}}(\theta )}{3\theta ^5}, \end{equation}

where $\mathcal{M}_{\textit{MVO}}(\theta )$ is given by (3.8). We note there are substantial differences for this coefficient in the presence of substrate adsorption: indeed, there is an increase in $A(\theta )$ relative to $A_{\textit{MVO}}(\theta )$ at the beginning of evaporation. This effect is reduced as $\mathcal{V}$ increases. As the droplet evaporates, the reduction of accumulated mass flux into the inner region from the droplet bulk leads to a substantial reduction (by several orders of magnitude at the later stages of evaporation) in $A(\theta )$ compared to (3.25).

Figure 3. The coefficient $A(\theta )$ for $\mathcal{V} = 0.25, 1, 4, 8$ (greyscales) and the equivalent coefficient in the no-substrate adsorption problem $A_{\textit{MVO}}(\theta )$ (blue). The dashed red curves represent the small-time and large-time limits (3.26) and (3.27).

These effects can be seen more clearly in the early- and late-time behaviour of the coefficient. In the early-time limit, we have

(3.26) \begin{align} A(\theta ) \sim \frac {8\pi \chi ^3}{\mathcal{V}\theta _0^{7/3}}\frac {3^{1/3}}{2^{7/6}}(\theta _0-\theta )^{1/3}\left [1-\frac {3^{1/3}\mathcal{V}}{2^{2/3}\theta _0^{1/3}}(\theta _0-\theta )^{1/3} + \frac {3^{2/3}V^2}{2^{7/3}\theta _0^{2/3}}(\theta _0-\theta )^{2/3}\right ]\\[-20pt]\nonumber \end{align}

as $\theta _0-\theta \rightarrow 0^+$ . In particular, we see that $A(\theta )\rightarrow 0$ as $\theta _0-\theta \rightarrow 0^+$ , retrieving the physically anticipated solution that the initial mass of colloid in the boundary layer is identically zero. Meanwhile, in the late-time limit, we find that

(3.27) \begin{equation} A(\theta ) \sim \frac {8\pi \chi ^3}{\mathcal{V}\theta _0^2}\left (\frac {e^{-\mathcal{V}\pi /6\sqrt {3}}}{3^{\mathcal{V}/2}\sqrt {2}}\right )\left (\frac {\theta }{\theta _0}\right )^{(\mathcal{V}-9)/4}\left [1 + \frac {(2\mathcal{V}-3)}{9}\left (\frac {\theta }{\theta _0}\right )^{3/4}\right ] \quad \mbox{as} \quad \theta \rightarrow 0^+. \end{equation}

These asymptotes are both indicated alongside the full solution for $A(\theta )$ in figure 3.

It is worth noting that the transition from $\mathcal{V} = 0$ to $\mathcal{V}\gt 0$ is not continuous in the coefficient $A(\theta )$ . Indeed, when $\mathcal{V} = O({\textit{Pe}}^{-1})$ as ${\textit{Pe}}\rightarrow \infty$ , the dominant balance in the matching procedure using global conservation of particle mass is different from that described above and we revisit this in § 3.6.

We note here that we can apply similar conservation of particle mass arguments for different evaporation models – specifically the one-sided model – as well, with similar findings in the reduction of the size of the nascent coffee ring. We discuss this in Appendix B.

3.3. Composite profiles

To view the particle distribution beyond the growing coffee ring, we must consider the droplet as a whole. However, we are unable to directly compute a leading-order composite profile for the volume fraction $\phi$ or the adsorbed mass $m_s$ due to their inverse square-root singularities at the contact line, which may not be matched to the exponentially decaying solutions in the boundary layer.

To circumvent the algebraic complexity of proceeding to higher order, we may follow Moore et al. (Reference Moore, Vella and Oliver2022) and introduce an intermediate region using the scalings

(3.28) \begin{equation} r = 1 - \frac {\mathring {r}}{{\textit{Pe}}^{k}}, \quad h = \frac {\mathring {H}}{{\textit{Pe}}^{k}}, \quad \bar {u} = {\textit{Pe}}^{k/2}\mathring {\bar {u}}, \quad \phi = {\textit{Pe}}\mathring {\phi }, \quad m_s = {\textit{Pe}}\mathring {m}_s \end{equation}

in (2.33)–(2.36), where $k\in (1,2)$ is chosen to be smaller than the contact line region, but large enough to enable the neglect of the time derivative terms in the advection–diffusion equation (2.33).

Under these scalings, retaining both advection and diffusion terms in (2.33), we find that

(3.29) \begin{equation} \frac {\partial }{\partial \mathring {r}}\left [\frac {2\chi \sqrt {\mathring {r}}}{{\textit{Pe}}^{k/2-1}\theta }\mathring {\phi } + \mathring {r}\frac {\partial \mathring {\phi }}{\partial \mathring {r}}\right ] = 0, \end{equation}

which yields the solution

(3.30) \begin{equation} \mathring {\phi }(\mathring {r},\theta ) = \mbox{e}^{-4\chi \sqrt {\mathring {r}}{\textit{Pe}}^{1-k/2}/\theta }\left [{\textit{Pe}} A(\theta ) + \frac {4\chi \gamma (\theta )}{\theta }\mbox{Ei}\left (\frac {4\chi \sqrt {\mathring {r}}{\textit{Pe}}^{1-k/2}}{\theta }\right )\right ]\!, \end{equation}

where $\mathrm{Ei}(x)$ is the exponential integral and the constants of integration have been chosen to match to the inner solution (3.16) as $\mathring {r}\rightarrow 0$ and the local expansion of the outer solution (3.5) as $\mathring {r}\rightarrow \infty$ . The corresponding solution for the intermediate adsorbed mass is given by

(3.31) \begin{equation} \mathring {m}_s(\mathring {r},t) = \frac {\mathcal{V}}{2\pi \theta _0}\int ^{\theta _0}_{\theta }\mathring {\phi }(\mathring {r},\psi )\,{\mathrm{d}} \psi . \end{equation}

We can construct a leading-order additive composite solution for the volume fraction using van Dyke’s rule:

(3.32) \begin{align} \displaystyle \phi _{\textit{comp}}(r,\theta ) & = \displaystyle \phi (r,\theta ) + {\textit{Pe}}\mathring {\phi }(r,\theta ) + {\textit{Pe}}^{2}\varPhi (r,\theta ) - \frac {\gamma (\theta )}{\sqrt {1-r}} \displaystyle - \mbox{exp}\left (-\frac {4\chi {\textit{Pe}}\sqrt {(1-r)}}{\theta }\right ) \nonumber\\ &\quad\times \left [{\textit{Pe}}^{2}A(\theta ) + \frac {4\chi \gamma (\theta )}{\theta }\log \left (\frac {4\chi {\textit{Pe}}\sqrt {(1-r)}}{\theta }\right )\right ]\!, \end{align}

where $\phi$ , $\mathring {\phi }$ and $\varPhi$ are the leading-order outer, intermediate and inner solutions given by (3.3)–(3.4), (3.30) and (3.16), respectively; the final term on the first line is the overlap contribution between the outer and intermediate regions; and the terms on the second line account for the overlap between the intermediate and inner regions, and the logarithmic singularity in the exponential integral, respectively.

Similarly, a leading-order additive composite solution for the adsorbed mass is given by

(3.33) \begin{align}& \displaystyle m_{s,\textit{comp}}(r,\theta ) = \displaystyle m_{s}(r,\theta ) + {\textit{Pe}}\mathring {m}_{s}(r,\theta ) + {\textit{Pe}}^{2}M_{s}(r,\theta ) - \frac {\mathcal{V}\mathcal{M}(\theta )}{\pi \chi \sqrt {1-r}} \displaystyle - {\textit{Pe}}\frac {\mathcal{V}}{2\pi \theta _0} \nonumber\\&\quad \times \int _{\theta }^{\theta _0} \mbox{exp}\left (\!-\frac {4\chi {\textit{Pe}}\sqrt {(1-r)}}{\psi }\right )\!\left [\!{\textit{Pe}} A(\psi ) + \frac {4\chi \gamma (\psi )}{\psi }\log \!\left (\!\frac {4\chi {\textit{Pe}}\sqrt {(1-r)}}{\psi }\right )\!\right ]{\mathrm{d}}\psi , \end{align}

where $m_{s}$ , $\mathring {m}_{s}$ and $M_{s}$ are the leading-order outer, intermediate and inner solutions given by (3.11), (3.31) and (3.17), respectively.

In figure 4, we display the composite mass profiles $m_{\textit{comp}} = h\phi _{\textit{comp}}$ , $m_{s,{comp}}$ alongside numerical results from the full advection–diffusion model (2.33)–(2.37) for ${\textit{Pe}} = 219,$ ${\textit{Da}} = 154$ ( $\mathcal{V} \approx 1.1$ ) based on parameters taken from Hu & Larson (Reference Hu and Larson2002) and Adamczyk et al. (Reference Adamczyk, Siwek, Weroński and Jaszczółt2003) as discussed in § 2.2. As the evaporation time increases, we see that the suspension mass in both the droplet bulk and the contact-line region decreases, as more and more particles are adsorbed onto the substrate, leading to an increase in the deposited mass throughout the droplet. Nevertheless, we see that the nascent coffee-ring profile persists in the suspension mass throughout the process, albeit with a gradually decreasing peak with advancing peak location, as illustrated in figure 4(b). This movement of the peak results in notable kinks in the deposited mass profile curves near the contact line, as illustrated in figure 4(d).

Figure 4. The leading-order composite suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 219$ , ${\textit{Da}} = 154$ ( $\mathcal{V} \approx 1.1$ ). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. The red dashed curves indicate the corresponding numerical solution of (2.33)–(2.37), while the arrows indicate increasing evaporation time.

The deposited mass is relatively uniform in the bulk of the droplet at all shown drying times, only substantially rising as we approach the contact line. The profile is monotonically increasing in $r$ : this mirrors the volume fraction, which is also monotonically increasing. Note that this is an inherent feature of the model herein: given that the flow velocity is increasing with $r$ , the depth-averaged thin-film model necessarily produces a monotonically increasing $\phi$ in the limit ${\textit{Pe}}\rightarrow \infty$ . This may not be the case if finite thickness effects are fully accounted for – indeed the analysis of Widjaja & Harris (Reference Widjaja and Harris2008) suggests both monotonic and non-monotonic profiles for the deposited mass – but this is beyond the scope of the present analysis.

We note that we see excellent agreement between the leading-order composite profiles and the numerical solutions throughout the evaporation time, which acts as a check on the veracity of our asymptotic results. Given that it is challenging to obtain good accuracy in the narrow boundary-layer region numerically, this validation gives us reassurance in using the asymptotic solution in this region.

In figure 4, we can clearly see that the dominant contributions to the leading-order distribution of particulate mass come from the outer suspension mass and the outer and inner deposited mass, with the size of the inner suspension mass significantly smaller. This is a vindication of the matching procedure used to obtain $A(\theta )$ and illustrates the departure from the previous theory in the absence of substrate adsorption (Moore et al. Reference Moore, Vella and Oliver2021).

3.4. Distribution of the mass

We may use our analysis to map the distribution of particle mass in the system throughout the drying process. A first metric to illustrate where the mass is at a given time is to calculate the incomplete integral mass functions, denoted by $G(r,\theta )$ and $H(r,\theta )$ for the suspension and deposited mass, respectively, where

(3.34) \begin{equation} G(r,\theta ) = 2\pi \int _{0}^{r} \bar {r}(h\phi )(\bar {r},\theta ) \,{\mathrm{d}}\bar {r}, \quad H(r,\theta ) = 2\pi \int _{0}^{r} \bar {r}m_{s}(\bar {r},\theta ) \,{\mathrm{d}}\bar {r}. \end{equation}

By construction, we initially have $G(r,\theta _0) = 1/2\pi$ , $H(r,\theta _0) = 0$ , while for $\theta \lt \theta _0$ , the mass redistributes itself according to the competing transport mechanisms of advection, diffusion and substrate adsorption. Here, we use the asymptotic results to approximate the integral masses to leading order as ${\textit{Pe}}\rightarrow \infty$ by using the composite solutions (3.32) and (3.33) in (3.34).

In figure 5, we display results for two specific cases: ${\textit{Pe}} = 219,\ {\textit{Da}} = 154$ ( $\mathcal{V} \approx 1.1$ ) and ${\textit{Pe}} = 40,\ {\textit{Da}} = 60$ ( $\mathcal{V} \approx 2.4$ ). The curves in the figures are coded by colour: the suspension mass is in orange/yellow, while the deposited mass is in blue. The curves are shown at $30\,\%$ , $60\,\%$ and $90\,\%$ of the drying time in each case. We see from the figure that when the relative effect of substrate adsorption is stronger, that is, with higher $\mathcal{V}$ , the proportion of particle mass deposited on the substrate grows more rapidly. Indeed, at $90\,\%$ of the drying time, there is ${\approx} 84\,\%$ of the total mass in the deposited layer for $\mathcal{V} \approx 1.1$ , while for $\mathcal{V}\approx 2.4$ , this rises to ${\approx} 97\,\%$ of the total mass.

Figure 5. The cumulative mass distribution in the suspension (orange-scale curves) and deposited onto the substrate (blue-scale curves) for (a) ${\textit{Pe}} = 219,\ {\textit{Da}} = 154$ ( $\mathcal{V} \approx 1.1$ ) and (b) ${\textit{Pe}} = 40,\ {\textit{Da}} = 60$ ( $\mathcal{V} \approx 2.4$ ). Results are displayed at $30\,\%,\ 60\,\%$ and $90\,\%$ of the drying time.

It is noticeable that for each case and across each time snapshot displayed, the majority of the integral suspension mass is congregated in the bulk of the droplet, rather than the boundary layer. As an example, for $\mathcal{V} \approx 1.1$ at $60\,\%$ of the drying time, the cumulative suspension mass at $1-r = 10^{-1}$ is ${\approx} 40\,\%$ of the total mass, while there is only ${\approx} 2\,\%$ more of the total mass between $1-r = 10^{-1}$ and $1-r = 10^{-6}$ . This is indicative of the relatively small amount of particle mass in the inner region: the presence of substrate adsorption reducing this from an $O(1)$ amount as described in Moore et al. (Reference Moore, Vella and Oliver2021) to an $O({\textit{Pe}}^{-2})$ amount here.

In contrast, as anticipated by the matching procedure in the asymptotics, the percentage of the total mass in the adsorbed layer close to the contact line is substantial: for the same example at the same dryout time, the cumulative deposited mass at $1-r = 10^{-1}$ is ${\approx} 28\,\%$ of the total mass, while between $1-r = 10^{-1}$ and $1-r = 10^{-6}$ , there is ${\approx} 29\,\%$ of the total mass. We are clearly embedded in a regime where the mass is primarily distributed between the suspension in the bulk, the deposited mass in the bulk and the deposited mass in the boundary layer. Notably, this changes over time, with the percentage left in suspension reducing as we approach dryout since the mass is necessarily deposited on the substrate according to the model.

We can investigate the mass distribution in the droplet bulk and local to the contact line as the droplet dries in more detail by considering each contribution separately. To do so, we choose $r = 1-1/{\textit{Pe}}$ as the boundary between the bulk and the contact line; this value is of course flexible, but we have selected it as it sits in the overlap between the outer and inner regions of the asymptotics. We then evaluate four integrals,

(3.35) \begin{align} \mathcal{G}_{{b}}(\theta ) &= 2\pi \!\!\int _{0}^{1-1/{\textit{Pe}}} \!r(h\phi _{\textit{comp}})(r,\theta )\, {\mathrm{d}} r, \quad\! \mathcal{G}_{{cl}}(\theta ) = 2\pi \!\!\int _{1-1\!/\!{\textit{Pe}}}^1 \!r(h\phi _{\textit{comp}})(r,\theta )\, {\mathrm{d}} r, \end{align}
(3.36) \begin{align} \mathcal{H}_{{b}}(\theta ) &= 2\pi \!\!\int _{0}^{1-1/{\textit{Pe}}} \!{\textit{rm}}_{s,{comp}}(r,\theta )\, {\mathrm{d}} r, \quad\! \mathcal{H}_{{cl}}(\theta ) = 2\pi \!\!\int _{1-1\!/\!{\textit{Pe}}}^1 \!{\textit{rm}}_{s,{comp}}(r,\theta )\, {\mathrm{d}} r, \end{align}

representing the suspension mass in the bulk, the suspension mass in the local contact-line region, the deposited mass in the bulk and the deposited mass in the contact-line region, respectively. We then demonstrate the role of substrate adsorption by fixing ${\textit{Pe}} = 40$ and varying ${\textit{Da}}$ so that we span from small $\mathcal{V}$ to large and show the results in figure 6.

Figure 6. Percentages of the total mass in suspension (black) and deposited (red) in the droplet bulk and in suspension (blue) and deposited (magenta) near the contact line as a function of drying time. In each case, ${\textit{Pe}} = 40$ and we take ${\textit{Da}} =$ (a) $10$ ( $\mathcal{V} \approx 0.39$ ), (b) $20$ ( $\mathcal{V} \approx 0.79$ ), (c) $60$ ( $\mathcal{V} \approx 2.36$ ) and (d) $100$ ( $\mathcal{V} \approx 3.92$ ).

The results largely corroborate our previous findings, particularly that the percentage of the total mass in the local region in suspension is substantially smaller than the other contributions, even for the relatively small $\mathcal{V} \approx 0.39$ example shown in figure 6(a). Of note is that the relative dominance of the deposited mass in the bulk and in the contact-line region changes with $\mathcal{V}$ : a stronger $\mathcal{V}$ favours deposition in the bulk as substrate adsorption begins to dominate advection. Even for $\mathcal{V}\approx 0.79 \lt 1$ , we find that at early stages the bulk deposited mass accounts for a greater percentage of the total mass than that in the contact-line region, although this switches after about $45\,\%$ of the drying time. In addition, we find that the evolutions of the deposited mass both in the bulk and in the contact-line region are concave up when substrate adsorption is relatively weak, as illustrated in figure 6(a,b), and are concave down when substrate adsorption dominates, as illustrated in figure 6(c,d). This indicates that more particles are deposited onto the substrate at early stages of evaporation when substrate adsorption is sufficiently strong, as expected.

We can further visualise the effect of $\mathcal{V}$ on the final distribution of mass by considering the residue densities – that is, adsorbed mass per unit area – in the droplet bulk and contact-line regions, defined by

(3.37) \begin{equation} \mathcal{D}_{{b}} = \frac {\mathcal{H}_{{b}}(0)}{\pi \big(1-{\textit{Pe}}^{-1}\big)^2}, \quad \mathcal{D}_{{cl}} = \frac {\mathcal{H}_{{cl}}(0)}{\pi \big(1-\big(1-{\textit{Pe}}^{-1}\big)^2\big)}, \end{equation}

respectively. The variation of these densities with $\mathcal{V}$ for different Péclet numbers is displayed in figure 7. Clearly, for smaller $\mathcal{V}$ , the dominance of advection in particle transport leaves a substantially larger density of material in the contact-line region: for example, for $\mathcal{V} = 1,\ {\textit{Pe}} = 40$ , we have $\mathcal{D}_{{cl}}/\mathcal{D}_{{b}} \approx 16.1$ . However, as the strength of substrate adsorption is increased through $\mathcal{V}$ , we see that there is a crossover value, beyond which the substantial deposition in the droplet bulk leads to a final density that is higher in the interior than at the contact line. In this sense, substrate adsorption is reducing the coffee-ring effect. For example, for ${\textit{Pe}} = 40$ this crossover point is $\mathcal{V} \approx 5.65$ . The crossover value clearly increases with ${\textit{Pe}}$ , as larger ${\textit{Pe}}$ correlates to stronger advection. Notably, across the entire range of Péclet numbers shown here, the density in the interior of the droplet remains largely unchanged: the main effect of the Péclet number is to increase the density in the contact-line region, again connected to the stronger coffee-ring effect for stronger advection.

Figure 7. The adsorbed mass per unit area in the droplet bulk $\mathcal{D}_{{b}}$ (solid curves) and contact-line region $\mathcal{D}_{{cl}}$ (dashed curves) for ${\textit{Pe}} = 20$ (black), $40, 60, 80, 100$ (light grey).

It is important to stress that the above comments referencing the effect of substrate adsorption on reducing the coffee-ring effect do not mean that there is less pointwise mass deposited near the contact line than in the bulk: indeed, since the volume fraction is always monotonically increasing in $r$ to leading order as ${\textit{Pe}}\rightarrow \infty$ , the same is true for the deposited mass, so that there is still an effective strengthening of the deposit at the contact line. This assumption will break down in regimes in which $\mathcal{V} \gg 1$ , since the dominant time scale is no longer that of the evaporation-driven flow, rather that dictated by substrate adsorption. This regime is beyond the scope of the present study.

3.4.1. The nascent coffee ring

To strengthen the above point, here it is pertinent to investigate how substrate adsorption influences the growth of the coffee ring. Since we are only interested in contributions near the contact line, we can forgo the composite profile in favour of the leading-order inner solutions. In previous studies in the absence of substrate adsorption, the nascent coffee ring has been identified, to leading order as ${\textit{Pe}}\rightarrow \infty$ , with the suspension mass in the inner region (Moore et al. Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022), which here is given by

(3.38) \begin{equation} m_{\textit{susp}}(R,\theta ) = H\kern-0.5pt\varPhi = \theta A\kern-0.5pt(\theta ) R\exp\left (-\frac {4\chi }{\theta }\sqrt {R}\right )\!. \end{equation}

However, in the presence of substrate adsorption, we must also consider the particle mass that has deposited on the substrate, which to leading order is simply

(3.39) \begin{equation} m_{\textit{dep}}(R,\theta ) = {\textit{Pe}}^2 M_{s} = {\textit{Pe}}^2 \frac {\mathcal{V}}{2\pi \theta _0}\int _{\theta }^{\theta _0} A(\psi )\exp\left (-\frac {4\chi }{\psi }\sqrt {R}\right )\,{\mathrm{d}} \psi . \end{equation}

If we interpret the coffee ring as the increase of particle mass due to evaporative transport to the contact line, we must consider the behaviour of both of these contributions.

Figure 8. The leading-order suspension (left) and deposited (right) mass in the inner region as a function of $R = {\textit{Pe}}^2(1-r)$ for $(a)$ $\mathcal{V}=0.25$ , $(b)$ $\mathcal{V}=1$ , $(c)$ $\mathcal{V}=4$ and $(d)$ $\mathcal{V}=8$ . In each case, we show profiles at $25\,\%$ $(\theta = 0.75\theta _0)$ , $50\,\%$ $(\theta = 0.5\theta _0)$ and $75\,\%$ $(\theta = 0.25\theta _0)$ of drying time: drying time is indicated by the arrow in each figure.

We display $m_{\textit{susp}}$ and $m_{\textit{dep}}$ as functions of $R$ for ${\textit{Pe}} = 10$ and $\mathcal{V} = 0.25,\ 1,\ 4,\ 8$ in figure 8. The curves are shown at $25\,\%$ , $50\,\%$ and $75\,\%$ of drying time in each case. In the figure, we see that the suspension mass retains its characteristic gamma distribution profile as discussed in detail in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022). However, not only is the size of this profile substantially reduced as compared with the zero-substrate adsorption regime, it also reduces with evaporation time, due to the increasing amount of particle mass lost to substrate adsorption. Notably, there is a simultaneous reduction of adsorbed mass in the inner region, which is due to the decrease in the coefficient $A(\theta )$ as $\mathcal{V}$ increases. Of course, these simultaneous local reductions happen concurrently with a rise in the particle mass deposited in the droplet bulk. This is consistent with a transition from advection-dominated particle transport to substrate adsorption-dominated particle transport as $\mathcal{V}\rightarrow \infty$ .

3.5. Breakdown of the model

While there are several limitations of the model we have presented, there are two particular potential points to investigate in terms of the breakdown of the theory, namely finite particle size effects in the suspension – sometimes referred to as jamming – and the growth of the deposited layer on the solid substrate. We use our asymptotic solution to discuss each of these in turn.

3.5.1. Finite particle size effects and jamming

Throughout, we have assumed that the particles are dilute in the fluid so that the flow and transport problems decouple. As we have seen, the largest volume fractions in the resulting flow are in the inner region in which ${\textit{Pe}}^2 \varPhi$ is monotonically increasing, with maximum value

(3.40) \begin{equation} \varPhi _{\textit{max}}(\theta ) = \varPhi (0,\theta )= A(\theta ), \end{equation}

where $A(\theta )$ is given by (3.24). Now, there will be a critical volume fraction $\tilde {\phi }_c$ at which the particles should no longer be considered as dilute, so that we expect to see a breakdown of the model at $\theta = \theta _{\textit{cb}}(\mathcal{V},{\textit{Pe}})$ such that

(3.41) \begin{equation} {\textit{Pe}}^2\varPhi _{\textit{max}}(\theta _{\textit{cb}})=\dfrac {\tilde {\phi }_c}{\tilde {\phi }_{\textit{init}}} \quad \implies \quad \frac {\gamma (\theta _{\textit{cb}})}{\theta _{\textit{cb}}^2} = \frac {\mathcal{V}\tilde {\phi }_c}{8\pi \chi ^3\tilde {\phi }_{\textit{init}}{\textit{Pe}}^2}. \end{equation}

We plot the curves of $\theta _{\textit{cb}}(\mathcal{V},{\textit{Pe}})$ as a percentage of the total drying time for varying $\mathcal{V}$ and ${\textit{Pe}} = 5,\ 10,\ 20,\ 40$ in figure 9 as the bluescale curves. We take the critical volume fraction to be $\tilde {\phi }_c = 0.64$ and for reference we take $\tilde {\phi }_{\textit{init}} = 0.05$ , based on the volume fraction for the data in table 4 in Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010), which, as discussed in § 2.2, looks at titania particles in water (calculated from mass fraction $X = 0.2$ in table 4, density of titania $\tilde {\rho }_p = 4230$ kg m $^{-3}$ , density of water $10^3$ kg m $^{-3}$ ).

Figure 9. The onset of jamming (bluescale curves, left-hand axis) and deposited layer effects (redscale curves, right-hand axis) as a percentage of the total drying time given by the roots of (3.41) and (3.46), respectively. For each case, we vary the Péclet number, ${\textit{Pe}} = 5,\ 10,\ 20,\ 40$ , while for $\theta _{{ct}}$ , we take $R = 1$ as an illustrative value.

For small substrate adsorption velocities, we observe similar behaviour to Moore et al. (Reference Moore, Vella and Oliver2021) in the sense that diffusive evaporation leads to rapid breakdown near the contact line. For example, for $\mathcal{V} = 0.4$ and ${\textit{Pe}} = 10$ , breakdown is predicted to occur after $0.002\,\%$ of the drying time for the values in the Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) experiments. However, it is worth again stressing that the presence of substrate adsorption here does inhibit finite particle size effects slightly (since volume fraction scales like ${\textit{Pe}}^2$ with substrate adsorption rather than ${\textit{Pe}}^4$ without).

When substrate adsorption is stronger, the onset of finite particle size effects is substantially delayed due to the particle mass lost to the substrate. For ${\textit{Pe}} = 10$ and $\mathcal{V} = 4$ , the suspension of particles may still be modelled as dilute for $88\,\%$ of the drying time. As can be seen, if we further increase $\mathcal{V}$ , there is a rapid decrease of the critical contact angle $\theta _{\textit{cb}}$ : studying the profiles of $A(\theta )$ in figure 3, a local maximum develops in $A$ near $\theta _0$ and for sufficiently large $\mathcal{V}$ , this maximum is below the jamming threshold, so that we only see jamming in the late stages when $A$ is singular as $\theta \rightarrow 0$ . This latter behaviour holds for $\mathcal{V} \lt 9$ (cf. (3.27)). For larger substrate adsorption, $A$ is bounded at $\theta = 0$ , so that jamming does not occur. However, as noted previously, for $\mathcal{V}\gg 1$ , we are no longer in the advection-dominated regime discussed herein.

3.5.2. Growth of the deposited layer

The second method of breakdown we discuss here is the invalidation of the assumption that we may neglect the deposited layer on the substrate in our analysis. This assumption is realised in the application of the no-slip, no-flux boundary conditions and the substrate adsorption condition on $z = 0$ , as well as the linear nature of the substrate adsorption condition. While this linearisation is reasonable as a first approximation initially, it may be called into question as the deposited layer grows. Not only may the size of the deposited layer play a role, but it may also cause a blocking effect, reducing the substrate adsorption velocity (Adamczyk Reference Adamczyk2002). Since the adsorbed mass is monotonically increasing in $r$ , this breakdown will necessarily occur first in the contact-line region, so that we may again concentrate our investigations using the solution in the inner region.

While, of course, the neglect of such a deposited layer will always be invalidated very close to the contact line, since the droplet free surface forms a wedge profile in the inner region, the modelling assumption has been that these local effects are small and do not propagate back towards the droplet bulk. However, clearly if the deposited layer becomes appreciable throughout the inner region, our analysis must be reconsidered.

We can estimate when this occurs by considering the leading-order deposited mass in the inner region. The dimensional mass deposited onto a region of radial extent $\tilde {\varDelta }$ about $\tilde {r} = \tilde {R}(1-R/{\textit{Pe}}^2)$ is given by

(3.42) \begin{equation} \tilde {m}_{s} \approx 2\pi \delta \tilde {R}^2 \tilde {\rho }_{{s}} \tilde {\phi }_{\textit{init}} {\textit{Pe}}^2 M_{s}(R,\theta ) \tilde {\varDelta }. \end{equation}

Now, if we suppose that the particles form a layer of thickness $\tilde {\epsilon }_{\textit{layer}}$ with packing fraction $0.64$ , we must have

(3.43) \begin{equation} \tilde {m}_{\textit{layer}} \approx 2\pi \tilde {R} \times 0.64\tilde {\rho }_{{s}} \tilde {\varDelta } \tilde {\epsilon }_{\textit{layer}} \approx 2\pi \delta \tilde {R}^2 \tilde {\rho }_{{s}}\tilde {\phi }_{\textit{init}} {\textit{Pe}}^2 M_{s}(R,\theta ) \tilde {\varDelta }. \end{equation}

The layer thickness can thus be estimated by

(3.44) \begin{equation} \tilde {\epsilon }_{\textit{layer}} \approx \frac {\delta \tilde {R} \tilde {\phi }_{\textit{init}}{\textit{Pe}}^2}{0.64} M_{s}(R,\theta ). \end{equation}

Meanwhile, the thickness of the droplet in this region is given by

(3.45) \begin{equation} \tilde {h} \approx \delta \tilde {R} {\textit{Pe}}^{-2}\theta R. \end{equation}

For the sake of comparison, let us suppose that our theory is invalidated if the deposited layer is 10 $\,\%$ of the droplet thickness: this occurs at the critical time $\theta = \theta _{{ct}}(\mathcal{V},{\textit{Pe}},R)$ , which is the root of

(3.46) \begin{equation} \frac {\mathcal{V}}{2\pi \theta _0}\int _{\theta _{{ct}}}^{\theta _0} A(\theta )\mbox{exp}\left (-\frac {4\chi }{\theta }\sqrt {R}\right )\,{\mathrm{d}} \theta = \frac {0.064}{\tilde {\phi }_{\textit{init}}{\textit{Pe}}^4} \theta _{{ct}} R. \end{equation}

Clearly, and as expected, the critical time for finite layer effects additionally depends on the distance from the contact line in the inner variables, $R$ . For illustration, we display in figure 9 as the redscale curves the breakdown time at $R = 1$ for ${\textit{Pe}} = 5,\ 10,\ 20,\ 40$ and varying $\mathcal{V}$ with $\tilde {\phi }_{\textit{init}} = 0.05$ taken from the experiments in Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010). Perhaps unsurprisingly, the factor of ${\textit{Pe}}^4$ in the denominator of (3.46) means that breakdown due to the growth of the deposited layer occurs rapidly as ${\textit{Pe}}\rightarrow \infty$ for all $\mathcal{V}$ , although since the initial adsorbed mass is zero, there is a timeframe of validity.

In addition, we see that the breakdown due to the finite layer thickness depends strongly on $\mathcal{V}$ . When we increase $\mathcal{V}$ , we increase the importance of substrate adsorption throughout the droplet, which reduces the amount adsorbed onto the substrate in the inner region, reducing the layer growth there and increasing the range of validity. For example, for ${\textit{Pe}} = 10$ , when $\mathcal{V} = 1$ , we see breakdown after ${\approx} 2.51\,\%$ of the total drying time, while for $\mathcal{V} = 10$ , this increases to ${\approx} 50.01\,\%$ .

It is also clear from figure 9 that for a fixed Péclet number and radial location $R$ , there is a transition from breakdown due to finite particle size effects to breakdown due to the deposited layer thickness as we increase $\mathcal{V}$ as the increasing significance of substrate adsorption results in more particles being found in the deposited layer rather than in suspension. In particular, for small $\mathcal{V}$ , the first breakdown is due to finite particle size effects. However, as we increase $\mathcal{V}$ , the reduction in the coefficient $A(\theta )$ coupled with the rapid growth of the deposited layer eventually leads to breakdown due to the latter. For example, if ${\textit{Pe}} = 10$ and $R = 1$ , this transition occurs at $\mathcal{V} \approx 1.52$ . With $R$ fixed, this transition value increases with ${\textit{Pe}}$ . Similarly, with ${\textit{Pe}}$ fixed, the transition value increases with $R$ .

However, for very small $\mathcal{V}$ , we note here an important caveat regarding the above results and discussion. As we noticed previously, the coefficient $A(\theta )$ and the matching procedure outlined in § 3.2.1 are no longer valid as $\mathcal{V}\rightarrow 0$ , due to the singular behaviour of the asymptotic expansion. We now present the resolution of this singularity.

3.6. The limit $\mathcal{V}\rightarrow 0$

When $\mathcal{V}$ is small, the asymptotic analysis presented previously breaks down when considering conservation of particle mass. For completeness, we briefly summarise the asymptotic results for $\mathcal{V} = o(1)$ as ${\textit{Pe}}\rightarrow \infty$ .

3.6.1. Outer region

In the outer region, substrate adsorption now plays a lower-order role in particulate transport so that to leading order as ${\textit{Pe}}\rightarrow \infty$ , the solution is precisely that presented in the no-substrate adsorption problem as given in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022):

(3.47) \begin{equation} \phi (r,\theta ) = \frac {1}{\sqrt {1-r^2}}\left (\left (\frac {\theta _0}{\theta }\right )^{3/4}-(1-(1-r^2)^{3/2})\right )^{1/3}\!. \end{equation}

As previously stated, this is regular in the sense that this is the leading-order solution if we take the limit of (3.3)–(3.4) as $\mathcal{V}\rightarrow 0$ . Note that the volume fraction is still square-root singular at the contact line: that is, $\phi \sim \gamma (\theta )/\sqrt {1-r}$ as $1-r\rightarrow 0^+$ , where

(3.48) \begin{equation} \gamma (\theta ) = \frac {1}{\sqrt {2}} \left (\left (\frac {\theta _0}{\theta }\right )^{3/4}-1\right )^{1/3} \end{equation}

is the leading-order term of (3.6) as $\mathcal{V}\rightarrow 0$ .

The corresponding leading-order solution for the deposited mass is given by $m_{s} \sim \mathcal{V}m_{s,0}$ , where

(3.49) \begin{align} m_{s,0}(r,\theta ) &= \frac {1}{2\pi \sqrt {1-r^2}(1-(1-r^2)^{3/2})}\nonumber\\&\quad\times\left [\left (1-\left (\frac {\theta }{\theta _0}\right )^{3/4}(1-(1-r^2)^{3/2})\right )^{4/3}- (1-r^2)^2\right ]\!. \end{align}

As we approach the contact line, we therefore have

(3.50) \begin{equation} m_{s} \sim \frac {\mathcal{V}\mathcal{M}_{\textit{MVO}}(\theta )}{\pi \chi \sqrt {1-r}} \quad \mbox{as} \quad 1-r\rightarrow 0^+, \end{equation}

as previously.

3.6.2. Inner region

In the inner region, the scalings (3.13) remain the same in the regime in which $\mathcal{V}\rightarrow 0$ , with the exception that we now have

(3.51) \begin{equation} \phi = \varphi \varPhi , \quad m_{s} = \mathcal{V}\varphi M_{s}, \end{equation}

where $\varphi$ is to be determined shortly. The solution then follows identically, with the leading-order inner particle volume fraction given by (3.16) and the leading-order inner deposited mass given by (3.17) with $\mathcal{V} =1$ .

3.6.3. Matching using conservation of particle mass

It is in the matching procedure that things differ most substantially as $\mathcal{V}\rightarrow 0$ . With the updated scales for the deposited mass in each region, (3.21) now becomes

(3.52) \begin{align} \displaystyle \frac {1}{2\pi } & = \displaystyle \int _{0}^{1} (rh\phi )_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + \varphi {\textit{Pe}}^{-4} \int _{0}^{\infty }\varPhi\!H(R,\theta )\,{\mathrm{d}} R \nonumber\\ &\quad +\mathcal{V}\displaystyle \int _{0}^{1} ({\textit{rm}}_{s,0})(r,\theta ) \,{\mathrm{d}} r + \mathcal{V}\varphi {\textit{Pe}}^{-2} \int _{0}^{\infty } M_s(R,\theta )\,{\mathrm{d}} R. \end{align}

Clearly, in this regime, the leading-order deposited mass in the outer region – represented by the first integral in the second line – no longer enters the leading-order balance in the conservation law. Nevertheless, we must still balance the mass lost out of the outer region to leading order due to advection. There are three distinct regimes.

For weak substrate adsorption when ${\textit{Pe}}^{-2}\ll \mathcal{V}\ll 1$ , the contribution due to substrate adsorption in the inner region dominates the contribution from the suspension mass in the inner region, so that we choose

(3.53) \begin{equation} \varphi = \mathcal{V}^{-1}{\textit{Pe}}^2. \end{equation}

Proceeding, we find that the coefficient $A(\theta )$ is then given by (3.24) with $\mathcal{V} = 1$ . In this weak substrate adsorption regime, substrate adsorption only affects the suspension mass to leading order in the inner region. Notably, the coefficient $A(\theta )$ is different from that of the zero-substrate adsorption problem, due to the different dominant terms in the conservation of mass equation. Furthermore, we note that substrate adsorption reduces the scaling for the suspension mass from that in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022). This is expected to reduce the role of finite particle size effects in the suspension, as discussed above for the case $\mathcal{V} = O(1)$ .

For the border case whereby $\mathcal{V} = \mathcal{W}{\textit{Pe}}^{-2}$ where $\mathcal{W} = O(1)$ as ${\textit{Pe}}\rightarrow \infty$ , the leading-order contributions from both the suspension mass and deposited mass in the inner region are of the same order of magnitude, so we take

(3.54) \begin{equation} \varphi = {\textit{Pe}}^4. \end{equation}

We may then derive the following integral equation for $A(\theta )$ :

(3.55) \begin{equation} \mathcal{M}_{\textit{MVO}}(\theta ) = \frac {3\theta ^5}{64\chi ^4}A(\theta ) + \frac {\mathcal{W}}{16\pi \theta _0\chi ^2}\int _{\theta }^{\theta _0} \psi ^2 A(\psi )\,{\mathrm{d}}\psi . \end{equation}

This has solution

(3.56) \begin{eqnarray} A(\theta ) = -\frac {64\chi ^4}{3\theta ^5}\int _{\theta }^{\theta _0} {\mathcal{M}^\prime}_{\textit{MVO}}(\psi )\,\mbox{exp}\left (\frac {2\mathcal{W}\chi ^2}{3\pi \theta _0}\left (\frac {1}{\theta ^2} - \frac {1}{\psi ^2}\right )\right )\,{\mathrm{d}} \psi . \end{eqnarray}

In this border regime, substrate adsorption has a leading-order effect on the suspension mass in the inner region only, through the exponential in the integral of (3.56). Moreover, the scaling for the size of the suspension mass is the same as that in the zero-substrate adsorption model of Moore et al. (Reference Moore, Vella and Oliver2021).

Finally, when substrate adsorption is very weak so that $\mathcal{V} = o({\textit{Pe}}^{-2})$ , each leading-order deposition term in (3.52) is small, so that we take

(3.57) \begin{equation} \varphi = {\textit{Pe}}^{4}, \end{equation}

and retrieve the solution of Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022) with $A = A_{\textit{MVO}}(\theta )$ as given by (3.25). Clearly, in this regime, substrate adsorption has a lower-order effect on the advection–diffusion competition that drives the growth of the nascent coffee ring. In principle, this will manifest as a temporal boundary layer close to lifetime, during which the increase in concentration results in substrate adsorption becoming significant, although in practice the model would likely break down prior to this point as a result of finite particle size effects.

3.6.4. Validation and comparisons

As previously, one may find composite profiles for the volume fraction $\phi$ and deposited mass $m_{s}$ in each of the three regimes discussed above.

Here, we investigate the veracity of the asymptotic analysis by considering a specific example for which ${\textit{Pe}} = 40$ and ${\textit{Da}} = 5$ , so that $\mathcal{V} \approx 0.20$ and we may reasonably say that $\mathcal{V} = O({\textit{Pe}}^{-1/2})$ , so that we are in the weak substrate adsorption regime. The results are shown in figure 10, where the suspension mass (figure 10 a,b) and deposited mass (figure 10 c,d) are shown at $15\,\%$ intervals of the drying time using the same colour coding as in figure 4. We see good agreement between the leading-order composite profiles and the numerical simulations in each case, with expected magnitude of errors given the relatively small Péclet number and large $\mathcal{V}$ chosen for this example. The behaviour of the nascent coffee ring is of particular interest here, with the peak initially rising for around $30\,\%$ of the drying time (cf. figure 10 b) due to the early-stage influence of the competition between advection and diffusion as reported in previous analyses (Moore et al. Reference Moore, Vella and Oliver2021). However, after this, the peak of the suspension mass profile begins to fall as the effect of substrate adsorption increases, reducing the mass left in the suspension at later times.

Figure 10. The leading-order composite suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 40$ , ${\textit{Da}} = 5$ ( $\mathcal{V} \approx 0.2$ ). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. The red dashed curves indicate the corresponding numerical solution of (2.33)–(2.37), while the arrows indicate increasing evaporation time.

Similar to figure 4, we see that the deposited mass is relatively uniform in the droplet bulk, although at later times we begin to see notable variation in $r$ away from the contact line. If we interpret the deposited mass as representative of the stain left after evaporation, this may suggest a slightly smoother transition from the ring to the internal deposit.

The figure also further illustrates that the particulate mass is predominantly contained in the outer suspension mass and the inner deposited mass in this regime, which justifies the balance chosen in the matching above. In particular, even at $90\,\%$ of the drying time, the outer deposited mass is still roughly $40\,\%$ smaller than the outer suspension mass.

It is worth noting that, while the composite solution is valid to leading order throughout the droplet, the intermediate solution for the volume fraction has a correction that is $O(1)$ as ${\textit{Pe}}\rightarrow \infty$ . Hence, for relatively small ${\textit{Pe}}$ or large $\mathcal{V}$ , it can often be more beneficial to directly use the leading-order outer/inner solutions for investigating behaviours in the droplet interior or at the contact line, respectively.

4. Stronger diffusion, $\boldsymbol{\textit{Pe}} = \boldsymbol{O(1)}$

In this section, we briefly discuss the regime in which the particle diffusive flux is larger, so that ${\textit{Pe}} = O(1)$ . Although coffee-ring formation is predominantly associated with a large Péclet number, since some of the examples discussed in § 2.2 feature ${\textit{Pe}} \sim 6{-}8$ , we briefly investigate numerically this regime and, in particular, discuss the role of substrate adsorption.

In figure 11, we plot simulation results for the parameter values for the droplet taken from Hu & Larson (Reference Hu and Larson2002) and for BSA particles from Kurrat et al. (Reference Kurrat, Prenosil and Ramsden1997) as discussed in § 2.2, for which ${\textit{Pe}} = 6.72$ and ${\textit{Da}} = 1.03$ , yielding $\mathcal{V} = 0.24$ . We note that despite the relatively small value of the Péclet number, the suspension mass still exhibits a characteristic nascent coffee-ring profile near the contact line. However, it is noticeable that the effect is weak, with a comparable height to the suspension mass at the droplet centre. It is likely that the relatively weak substrate adsorption compared with advection in this example contributes to this sustained ring profile in the suspension.

Figure 11. Numerical results for the suspension (a,b) and deposited (c,d) mass as a function of time for data taken from Kurrat et al. (Reference Kurrat, Prenosil and Ramsden1997) and Hu & Larson (Reference Hu and Larson2002), for which ${\textit{Pe}} = 6.72$ , ${\textit{Da}} = 1.03$ ( $\mathcal{V} \approx 0.24$ ). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. Arrows indicate increasing evaporation time.

The smaller Péclet number also leads to the ring forming further from the contact line (cf. the scaling for the inner region in the asymptotic analysis). This leads to a flatter profile of the adsorbed mass in the contact-line region, with little variation for $1-r\lesssim 10^{-3}$ . In the bulk, the adsorbed mass profile is similar to that in figure 10, which is consistent with the similar values of $\mathcal{V}$ , although in the present example, the weaker advection allows for slightly more adsorbed mass in the droplet bulk.

We consider a droplet with even weaker advection in figure 12, where ${\textit{Pe}} = 1$ , ${\textit{Da}} = 0.5$ so that $\mathcal{V} \approx 0.79$ . One can clearly see the diminished coffee-ring effect, with no substantial buildup of suspension mass in the contact-line region until around $90\,\%$ of the drying time. This late-time emergence of the coffee-ring profile is due to the large time-dependent Péclet number ${\textit{Pe}}/(1-t)$ as $t\rightarrow 1^-$ , as discussed in detail in Moore et al. (Reference Moore, Vella and Oliver2021). However, we note that the amount of mass in the contact-line region in suspension remains very small.

Figure 12. Numerical results for the suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 1$ , ${\textit{Da}} = 0.5$ ( $\mathcal{V} \approx 0.79$ ). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. Arrows indicate increasing evaporation time.

Due to the reduced transport of particle mass to the contact line, the adsorbed mass profile is much more uniform than in previous examples, with a greater buildup of deposit in the droplet bulk and a reduced deposit near the contact line. Indeed, we see that at $90\,\%$ of the drying time, the ratio of the adsorbed mass at the contact line to that at the centre is ${\approx} 12.8$ , compared with a ratio of ${\approx} 1.95\times 10^{3}$ in the previous example displayed in figure 11. This suggests a much more uniform final deposit, as we would anticipate in this regime and consistent with previous studies (e.g. Widjaja & Harris Reference Widjaja and Harris2008).

Naturally, the reduced coffee-ring effect for ${\textit{Pe}} = O(1)$ reduces the possibility of breakdown of the model due to finite particle size effects, while the associated enhanced substrate adsorption increases breakdown due to the growth of the deposited layer.

5. Comparison with other models and experiments

We conclude by comparing the predictions of our model with other models in the literature and available experimental data. Given the complexity of the problem, it is perhaps unsurprising that finding suitable comparisons can be a challenge, but here we present three examples.

5.1. Comparison with previous numerical studies

First, we consider the simulations in Widjaja & Harris (Reference Widjaja and Harris2008), where the fully coupled diffusive evaporation, Stokes flow and particle transport problems are solved numerically. The majority of the examples studied therein are in the regime that ${\textit{Da}} \gg {\textit{Pe}}$ , and thus fall outside the scope of this paper. However, an example that falls into our scope is that in their figure 9(a), for which ${\textit{Pe}} = {\textit{Da}} = 12.5$ (where we have accounted for the slightly different definition of the Péclet and Damköhler numbers herein). Note that the droplet under consideration has a relatively large aspect ratio, $\delta = 0.8$ , which pushes the limits of applicability of the depth-averaged model we use.

Figure 13. A comparison between simulation data from figure 9(a) in Widjaja & Harris (Reference Widjaja and Harris2008) (circles), the leading-order composite asymptotic adsorbed mass given by (3.33) (solid grey curve) and the numerical solution of (2.33)–(2.37) (dashed red curve). In the language of the present paper, the results are presented for ${\textit{Da}} = {\textit{Pe}} = 12.5$ $(\mathcal{V} = \pi /2)$ .

In figure 13, we compare data extracted from figure 9(a) in Widjaja & Harris (Reference Widjaja and Harris2008) with the leading-order composite solution (3.33) and the numerical solution of (2.33)–(2.37). We see that, despite the relatively large initial droplet aspect ratio, the model presented herein does a good job of qualitatively capturing the monotonic increasing nature of the adsorbed mass profile. The agreement is particularly good in the droplet bulk, although our model slightly under-predicts the profile near the contact line. It is anticipated that this is a combination of the large aspect ratio and that the Péclet number is quite small, so that the assumption that $\delta ^2{\textit{Pe}} \ll 1$ made in deriving our depth-averaged model is not met well by this example. Nevertheless, despite being at the limits of the validity of the model, these comparisons are encouraging.

5.2. Comparison with previous experimental studies

Let us now turn to experiments. Generally, the qualitative behaviour that more mass is adsorbed onto the substrate in the bulk when there is an increase in substrate adsorption, e.g. when particle–substrate interaction is attractive rather than repulsive, is a common observation in experimental studies (see e.g. Yan et al. Reference Yan, Gao, Sharma, Chiang and Wong2008; Bhardwaj et al. Reference Bhardwaj, Fang, Somasundaran and Attinger2010; Anyfantakis & Baigl Reference Anyfantakis and Baigl2015; Devineau et al. Reference Devineau, Anyfantakis, Marichal, Kiger, Morel, Rudiuk and Baigl2016; Molchanov et al. Reference Molchanov, Roldughin, Chernova-Kharaeva and Senchikhin2019; Kumar et al. Reference Kumar, Basavaraj and Satapathy2023). However, as discussed previously in § 2.2, it is challenging to get experimental data that accurately estimates all the required parameters in the model to make a quantitative comparison, with the substrate adsorption velocity $\tilde {k}_{{a}}$ proving the most troublesome. In addition, due to the importance of finite particle size effects such as jamming or particle packing in the vicinity of the coffee ring and in the deposited layer, the model presented herein is unlikely to give good quantitative comparisons with the final deposited profile. However, given that for the vast majority of problems, the dilute regime is the natural starting point – and that we further demonstrated that substrate adsorption can extend the validity of this regime – we are more likely to have success when comparing quantities such as the mass in the ring and the central deposit. Here we carry out a quantitative comparison with two experimental studies by Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) and Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020), in which water droplets containing silica and titania nanoparticles evaporate on gold-coated silicon wafer and glass substrates, respectively, for which we were able to determine approximations for $\tilde {k}_{{a}}$ as outlined in § 2.2. These studies also provide profiles of the final deposit from which we may determine the mass in the ring and the central deposit.

Before making our comparisons it is important to clarify that the experiments considered here study droplets evaporating with a pinned contact line for most of their lifetime. That is, the contact line will eventually de-pin towards the end of evaporation. In particular, for the experiments by Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) the de-pinning angle is approximately $ 6^\circ$ , indicating that the contact line is pinned for only $86\,\%$ and $84\,\%$ of lifetime for the two experiments of untreated and aluminium-coated silica nanoparticles, respectively, outlined in § 2.2. In addition, close inspection of supplementary videos 1 and 2 in Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) indicate that the droplet is pinned for ${\approx} 93\,\%$ of lifetime when $\text{pH}=11.7$ but only for ${\approx} 70\,\%$ of lifetime when $\text{pH}=2.8$ . Therefore, in order to test the validity of the present model we compare quantities for the mass in the ring when the contact line de-pins, i.e. at $\theta =\theta ^*$ , rather than at $\theta =0$ , assuming that after the contact line recedes no more mass can be deposited into the ring.

Figure 14 shows a comparison between the leading-order deposited mass in the contact-line region at the time of contact-line de-pinning $\mathcal{H}_{{cl}}(\theta ^*)$ with the experimental data for the final mass in the ring from Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) and Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020). Specifically, as figure 14 shows, despite the number of simplifying assumptions made herein, the asymptotic solution does a good job of predicting the final mass in the ring deposit, slightly over-predicting the experimental values of Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) by approximately $6\,\%$ and $9\,\%$ , but with particularly good agreement when comparing with the experiments of Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010). The latter is perhaps unsurprising given that this example is comfortably in the regime in which ${\textit{Pe}}\gg 1$ and $\delta \ll 1$ .

Figure 14. Comparison between the leading-order deposited mass in the contact-line region at the time of contact-line de-pinning $\mathcal{H}_{{cl}}(\theta ^*)$ given by (3.36) for (a) ${\textit{Pe}}\approx 50.0$ , ${\textit{Da}}\approx 3.66$ and $\mathcal{V}\approx 0.115$ (black line) and ${\textit{Pe}}\approx 69.1$ , ${\textit{Da}}\approx 4.47$ and $\mathcal{V}\approx 0.102$ (grey line) and the experimental data for the final mass in the ring from supplementary figures S8 and S9 in Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) for untreated (star) and aluminium-coated (diamond) particles, and for (b) ${\textit{Pe}}\approx 127$ , ${\textit{Da}}\approx 20.5$ and $\mathcal{V}\approx 0.253$ (black line) and ${\textit{Pe}}\approx 213$ , ${\textit{Da}}\approx 8.23$ and $\mathcal{V}\approx 0.061$ (grey line) and experimental data for the final mass in the ring from figures 6(a) and 6(b) in Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) for $\text{pH}=11.7$ (circle) and $\text{pH}=2.8$ (square). The dashed lines correspond to the predicted mass in the contact-line region when the droplet evaporates with a pinned contact line throughout evaporation $\mathcal{H}_{{cl}}(0)$ .

In all the cases considered, the predicted mass in the contact-line region at $\theta =0$ is substantially larger than the true value from experiments, as well as those predicted by the model at the time of contact-line de-pinning, highlighting the importance of contact-line motion in the formation of the final deposit. For example, for the experiments by Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) in which the droplets are pinned for $84\,\%$ $86\,\%$ of lifetime, the final predicted adsorbed mass in the contact-line region after contact-line de-pinning and in the bulk after complete evaporation is approximately $74\,\%$ and $8\,\%$ $9\,\%$ of the total mass, respectively. This indicates that approximately $17\,\%$ $18\,\%$ of the mass is left unaccounted for and therefore must be deposited in the bulk due to late contact-line de-pinning. This proportion of final mass is nearly double the predicted mass deposited in the bulk due to substrate adsorption effects. We note that whilst the majority of previous studies on droplet evolution and particle deposition often ignore late contact-line de-pinning, a recent study by D’Ambrosio et al. (Reference D’Ambrosio, Wray and Wilson2025a ) gives theoretical predictions for the final mass of particles deposited on the substrate split between the contact-line region and the bulk for a variety of de-pinning angles, $\theta ^*$ , for a thin droplet of dilute suspension in the absence of substrate adsorption. Specifically, their model predicts that for the experiments by Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020) approximately $23\,\%$ $25\,\%$ of the total mass will be deposited in the bulk due to contact-line de-pinning, which are similar values to those outlined here although slightly higher presumably due to the absence of substrate adsorption in their model.

Moreover, it is clear from the experiments by Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) that substrate adsorption has a strong effect on contact-line pinning. Specifically, for the experiment when $\text{pH}=11.7$ , the DLVO force is repulsive, substrate adsorption is weak and particles are advected to the contact line which remains pinned for ${\approx} 93\,\%$ of the evaporation. The predicted de-pinning angle of ${\approx} 0.24^\circ$ is significantly lower than the value predicted for clean water on glass by Hu & Larson (Reference Hu and Larson2002), namely $2^\circ$ $4^\circ$ , showing that the strong coffee-ring effect enhances the pinning in this case. However, when $\text{pH}=2.8$ , the DLVO force is attractive and fewer particles are advected to the contact line due to substrate adsorption, i.e. the coffee-ring effect is reduced. The predicted de-pinning angle of ${\approx} 1.96^\circ$ is much closer to the value for clean water and the earlier de-pinning time further reduces the coffee-ring effect in this case as more particles are deposited in the bulk as the contact line recedes.

6. Discussion and summary

In this study, we have investigated the competing effects of particle advection, diffusion and substrate adsorption in thin, pinned sessile droplets evaporating under a diffusive evaporative flux. Under the assumption that the capillary number of the droplet is small and that the particles remain dilute in the suspension, we derived a model for the particle distribution in the suspension and that adsorbed onto the substrate that depends on two critical parameters, the particle Péclet number ${\textit{Pe}}$ , which is the ratio of the advective and diffusive fluxes on the droplet scale, and the Damköhler number ${\textit{Da}}$ , which is the ratio of substrate adsorption to diffusion.

By referring to a number of experimental studies, we demonstrated that for a wide variety of real-world systems, the pertinent limit is that in which the Péclet number satisfies ${\textit{Pe}}\gg 1$ , so that particle advection dominates diffusion within the bulk of the droplet, with a boundary layer near the contact line where the effects are comparable and the nascent coffee ring grows. Further, we saw that while ${\textit{Da}}$ could take a range of values, we consistently have ${\textit{Da}}\lesssim {\textit{Pe}}$ . We therefore focused our attention on the regime in which the ratio between the substrate adsorption and diffusion particle velocities $\mathcal{V} = \pi {\textit{Da}}/(2 {\textit{Pe}}) = O(1)$ as ${\textit{Pe}}\rightarrow \infty$ .

Using these parameter estimates, we performed a detailed asymptotic analysis in the limit in which ${\textit{Pe}}\rightarrow \infty$ . Our analysis is an extension of that in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022) in the absence of substrate adsorption. In particular, the asymptotic structure consists of an outer problem in the bulk of the droplet in which particle advection and substrate adsorption are the dominant effects and the solution follows that of D’Ambrosio et al. (Reference D’Ambrosio, Wray and Wilson2025b ). Local to the contact line, the suspension particle volume fraction is singular, necessitating a boundary layer at the contact line in which particle advection and diffusion balance, with substrate adsorption a lower-order effect. While the inclusion of substrate adsorption does not alter the radial size of the boundary layer, which is $O({\textit{Pe}}^{-2})$ , the loss of particle mass in the bulk due to advection is now compensated by the mass deposited in the boundary layer, as opposed to that in suspension as seen in Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022). This correspondingly reduces the amount of mass in suspension from $O({\textit{Pe}}^2)$ to $O(1)$ in the presence of substrate adsorption. Perhaps surprisingly, this leading-order balance in the global solute mass persists for moderately small $\mathcal{V}$ : indeed for ${\textit{Pe}}^{-2}\ll \mathcal{V}\ll 1$ , we find that the suspension mass scales like $\mathcal{V}^{-1}$ in the boundary layer. It is only for very weak substrate adsorption $\mathcal{V} = O({\textit{Pe}}^{-2})$ that we retrieve the same scalings as those of the zero-substrate adsorption regime of Moore et al. (Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022).

The leading-order solution for the particle volume fraction in the boundary layer takes the same exponential form as that in the absence of substrate adsorption, with the exponent depending on the local contact angle $\theta$ , the coefficient of evaporative flux singularity at the contact line $\chi$ and the radial position $R = {\textit{Pe}}^{2}(1-r)$ . The role of substrate adsorption is to change the coefficient of this exponential, which again depends on $\theta$ and $\chi$ , but also on $\mathcal{V}$ and the derivative of the mass flux advected into the contact line, $\mathcal{M}'(\theta )$ .

We compared our asymptotic results with numerical simulations of the advection–diffusion equation, demonstrating excellent agreement, enabling us to infer the effect of substrate adsorption on the coffee-ring effect and the final distribution of particle mass. In particular, when $\mathcal{V} = O(1)$ as ${\textit{Pe}}\rightarrow \infty$ , we showed that substrate adsorption drastically reduces the total mass in the boundary-layer region by a factor of $O({\textit{Pe}}^{-2})$ compared with the zero-substrate adsorption regime. Moreover, at a fixed drying time, increasing $\mathcal{V}$ leads to a transition from the mass primarily being distributed between the bulk suspension and the adsorbed mass near the contact line to being distributed between the bulk suspension and the bulk adsorbed mass. This is due to the relative increase in the importance of the substrate adsorption velocity compared with advection.

It is worth stressing that for all cases considered in this paper, including those for which ${\textit{Pe}} = O(1)$ , the final deposited mass distribution still exhibits a maximum at the contact line, which may be interpreted as a coffee-ring effect, even in cases where advection is weaker. However, there is a clear reduction in the maximum and corresponding promotion of a more uniform deposit both with reduced ${\textit{Pe}}$ and with increased $\mathcal{V}$ .

We discussed two possible mechanisms for the breakdown of our model: the role of finite particle size effects and the growth of the deposited layer. In the former case, the assumption that the particle mass is dilute in the fluid decouples the flow and particle transport problems, simplifying the resulting analysis. However, in the limit that ${\textit{Pe}}\rightarrow \infty$ , the volume fraction increases local to the contact line. In the latter, we neglected the deposited layer within our model, which enabled us to linearise boundary conditions onto the initial substrate height and assume a linear substrate adsorption regime. However, particularly for stronger substrate adsorption, this may be invalidated, particularly close to the contact line, where the thickness of the droplet vanishes.

We explored the onset of breakdown for a range of different ${\textit{Pe}}$ and $\mathcal{V}$ . When the role of substrate adsorption is very small $\mathcal{V} = O({\textit{Pe}}^{-2})$ , the primary form of breakdown is due to finite particle size effects, since the volume fraction has the same boundary-layer scaling $O({\textit{Pe}}^{4})$ as for the zero-substrate adsorption regime discussed in Moore et al. (Reference Moore, Vella and Oliver2021). When ${\textit{Pe}}\gg 1$ , the critical volume fraction for jamming may be reached in very short time, necessitating the inclusion of a jammed region for the majority of the drying process. This effect is reduced as ${\textit{Pe}}$ decreases.

As we increase the importance of substrate adsorption by increasing $\mathcal{V}$ , we delay the onset of breakdown due to finite particle size effects, while simultaneously promoting breakdown due to the growing deposited layer. For a fixed Péclet number, this trend continues until a transition from the former to the latter occurs at a specific value of  $\mathcal{V}$ : in the regime where $\mathcal{V} = O(1)$ as ${\textit{Pe}}\rightarrow \infty$ , this transition occurs for larger $\mathcal{V}$ as ${\textit{Pe}}$ increases. Nevertheless, it is worth noting that there remains a significant part of the parameter space in which neither method of breakdown has triggered and the dilute suspension and negligible deposited layer thickness assumptions are reasonable, so that the predictions presented in our model may be readily used to discern the particle mass distribution.

We compared our leading-order asymptotic predictions with previous numerical and experimental studies. Specifically, we found encouraging agreement between the leading-order composite solution for the adsorbed mass and the numerical solution with data extracted from the simulations of Widjaja & Harris (Reference Widjaja and Harris2008), despite the relatively large initial aspect ratio of the droplet and low Péclet number in this example. In addition, we found good agreement between the predictions of the present model for the adsorbed mass in the contact-line region at the time of contact-line de-pinning and experimental data for the final mass in the ring from two different studies by Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) and Bridonneau et al. (Reference Bridonneau, Zhao, Battaglini, Mattana, Thévenet, Noël, Roché, Zrig and Carn2020). The agreement is particularly good when comparing with Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) in which the initial aspect ratio of the droplet and the Péclet number are comfortably in the regime $\delta \ll 1$ and ${\textit{Pe}}\gg 1$ . We note that the comparisons suggest that, at least for the experiments considered here, contact-line de-pinning can cause a more significant reduction in the ring deposit than substrate adsorption effects, even when the droplet is pinned for most ( $85\,\%$ $95\,\%$ ) of lifetime. Therefore, in order to facilitate comparisons between the mass in the contact-line region and the bulk at the end of evaporation, the present model would need to be extended to consider contact-line motion.

Another natural next step would be to consider an extension to the model that may include a growing jammed region of particles or the finite size of the deposited layer to extend the analysis further into the evaporative process. Certainly, in order to quantitatively compare pointwise with the final distribution patterns seen in experiments such as those of Bhardwaj et al. (Reference Bhardwaj, Fang, Somasundaran and Attinger2010) or Kajiya et al. (Reference Kajiya and Kaneko2008), a more careful inclusion of, in particular, finite particle size effects will be necessary. A natural first step may be to assume that the fluid is dilute up to a fixed jamming volume fraction as in Popov (Reference Popov2005). However, more complicated models including volume-fraction-dependent viscosities and diffusivities are also available (see e.g. Kaplan & Mahadevan Reference Kaplan and Mahadevan2015; Guazzelli & Pouliquen Reference Guazzelli and Pouliquen2018). Our analysis would inform the initial conditions before the onset of any such jamming model.

Further aspects of particle dynamics that we have not considered in our model include particle aggregation (Zhang et al. Reference Zhang, Sibley, Tseluiko and Archer2024) and the possibility of free-surface capture, which is known to be important for a wide variety of droplets (D’Ambrosio et al. Reference D’Ambrosio, Wilson, Wray and Duffy2023), particularly thicker droplets (Kang et al. Reference Kang, Vandadi, Felske and Masoud2016). There has also been a recent attempt to include a jammed surface layer in such cases (Coombs et al. Reference Coombs, Sprittles and Chubynsky2024b ), although this analysis neglects the importance of radial diffusive effects in particle transport and does not allow for substrate adsorption.

Acknowledgements

A.W.W. and M.R.M. would like to acknowledge the support of the Royal Society of Edinburgh via a Research Collaboration Grant and the Carnegie Trust for support via Research Incentive Grant RIG013348. Part of this work was supported by an INI In-Residence Programme and M.R.M. was partially supported by a grant from the Simons Foundation.

Declaration of interests

The authors report no conflict of interests.

Appendix A. Numerical solution

We wish to solve the evolution equations for $h\phi$ (2.33) and $m_{s}$ (2.36) subject to no flux conditions at the centreline and contact line (2.35) and a uniform initial concentration (2.34). This can be performed by a modification of the methodology of Moore et al. (Reference Moore, Vella and Oliver2021). In particular, we define the incomplete radial integral mass:

(A1) \begin{equation} \varGamma (r,t)=\int _0^r \bar {r}h(\bar {r},t)\phi (\bar {r},t){\,\mathrm{d}} \bar {r}. \end{equation}

In terms of this variable, (2.33) and (2.36) can be recast in the forms

(A2) \begin{align}& \frac {\partial \varGamma }{\partial t}+\frac {1}{4}\left [ \bar {u}+\frac {1}{r{\textit{Pe}}}+\frac {1}{{\textit{Pe}} h}\frac {\partial h}{\partial r} \right ]\frac {\partial \varGamma }{\partial r}-\frac {1}{4{\textit{Pe}}}\frac {\partial ^2\varGamma }{\partial r^2}+\frac {{\textit{Da}}}{4{\textit{Pe}}}\int _0^r \frac {1}{h}\frac {\partial \varGamma }{\partial r}{\,\mathrm{d}} r=0, \end{align}
(A3) \begin{align}&\qquad\qquad\qquad\qquad\qquad \frac {\partial m_s}{\partial t}=\frac {{\textit{Da}}}{4{\textit{Pe}}}\frac {1}{rh}\frac {\partial \varGamma }{\partial r}, \end{align}

where $\varGamma$ satisfies Dirichlet conditions at each end of the domain, namely

(A4) \begin{equation} \varGamma (0,t)=0, \quad \varGamma (1,t)=\int _{0}^1 rh\phi {\,\mathrm{d}} r. \end{equation}

The second condition here can be written in a more suitable form by considering the evaluation of (2.33) at $r=1$ , yielding

(A5) \begin{equation} \frac {\partial }{\partial t}\left (\int _0^1 rh\phi {\,\mathrm{d}} r\right )=-\frac {{\textit{Da}}}{4{\textit{Pe}}}\int _0^1 \frac {1}{h}\frac {\partial \varGamma }{\partial r}{\,\mathrm{d}} r=-\int _0^1 r\frac {\partial m_s}{\partial t}{\,\mathrm{d}} r. \end{equation}

Integrating this gives

(A6) \begin{equation} \varGamma (1,t)=\frac {1}{2\pi }-\int _{0}^1 r m_s {\,\mathrm{d}} r. \end{equation}

The initial condition is

(A7) \begin{align} \varGamma (r,0)=\int _{0}^r \bar {r}h(\bar {r},0)\phi (\bar {r},0)\,{\mathrm{d}}\bar {r}=\frac {2}{\pi }(1-t)(1-r^2)|_{t=0}=\frac {2}{\pi }\left (\frac {r^2}{2}-\frac {r^4}{4} \right )\!. \end{align}

This system is solved using a code that is second-order in space, using the same non-uniform grid spacing as described by Moore et al. (Reference Moore, Vella and Oliver2021), and first-order semi-implicit in time. In particular, the non-local contributions given by the final terms in (A2) and (A6) would introduce strong coupling, making the Jacobian matrices substantially less sparse, and much more expensive to solve. As a consequence, these terms are treated explicitly, and the rest implicitly. Both time step and grid size were refined to ensure convergence, and in general a time step of $\delta t=10^{-5}$ was used with $10^5$ grid points.

Appendix B. One-sided evaporation

In this appendix, we briefly discuss the role of substrate adsorption for a different evaporation model, namely kinetic or one-sided evaporation, in which the limiting factor for evaporation is the transport of liquid to the interface (Murisic & Kondic Reference Murisic and Kondic2011).

For thin droplets, it is a reasonable first model to take the resulting evaporative flux to be constant (Moore et al. Reference Moore, Vella and Oliver2021, Reference Moore, Vella and Oliver2022). The resulting evaporation-driven velocity is finite at the contact line, but a coffee ring persists nonetheless. While we do not go into the details of the full analysis in such a regime, we discuss the application of conservation of mass to determine the scaling for the size of the coffee ring (cf. § 3.2.1).

Let us focus on the case that $\mathcal{V}=O(1)$ as ${\textit{Pe}}\rightarrow \infty$ . In the outer region, both the particle volume fraction $\phi$ and adsorbed mass $m_s$ are therefore $O(1)$ as ${\textit{Pe}}\rightarrow \infty$ , while according to Moore et al. (Reference Moore, Vella and Oliver2021), the pertinent scalings for the inner region are

(B1) \begin{equation} r = 1 - \frac {R}{{\textit{Pe}}}, \quad h = \frac{H}{\textit{Pe}}, \quad \phi = {\textit{Pe}}^c \varPhi , \quad m_s = {\textit{Pe}}^c M_s, \end{equation}

where we note that the boundary layer is an order of magnitude larger than for diffusion-limited evaporation and $c$ is to be determined by matching.

Following the same argument as § 3.2.1, using conservation of particle mass, we have

(B2) \begin{equation} \displaystyle \frac {1}{2\pi } = \displaystyle \int _{0}^{1} (\textit{rh}\phi + {\textit{rm}}_s)_{\textit{outer}}(r,\theta ) \,{\mathrm{d}} r + {\textit{Pe}}^{c-2} \!\int _{0}^{\infty }\! {H\kern-0.5pt\varPhi (\!R,\theta )\,{\mathrm{d}} R} + {\textit{Pe}}^{c-1} \!\int _{0}^{\infty }\! M_s(\!R,\theta )\,{\mathrm{d}} R. \end{equation}

Clearly, to account for the mass lost into the inner region, we must take $c = 1$ ; this is an order of magnitude smaller than the equivalent problem in the absence of substrate adsorption (i.e. when $\mathcal{V} = 0$ ). We thus see a similar reduction in the size of the nascent coffee ring for one-sided evaporation as well.

Similar logic to § 3.6.3 also applies when $\mathcal{V} = o(1)$ as ${\textit{Pe}}\rightarrow \infty$ , with the switch over from the dominance of particle advection and diffusion to substrate adsorption, advection and diffusion occurring when $\mathcal{V} = O({\textit{Pe}}^{-1})$ as ${\textit{Pe}}\rightarrow \infty$ .

References

Abo Jabal, M., Homede, E., Pismen, L.M., Haick, H. & Leshansky, A.M. 2017 Controlling Marangoni flow directionality: patterning nano-materials using sessile and sliding volatile droplets. Eur. Phys. J. Special Topics 226 (6), 13071324.10.1140/epjst/e2016-60404-xCrossRefGoogle Scholar
Adamczyk, Z. 2002 Irreversible adsorption of particles. In Adsorption: Theory, Modeling and Analysis (ed. T. József), pp. 251374. M. Dekker.Google Scholar
Adamczyk, Z. 2019 Protein adsorption: a quest for a universal mechanism. Curr. Opin. Colloid Interface Sci. 41, 5065.10.1016/j.cocis.2018.11.004CrossRefGoogle Scholar
Adamczyk, Z., Dabros, T., Czarnecki, J. & Van De Ven, T.G.M. 1983 Particle transfer to solid surfaces. Adv. Colloid Interface Sci. 19 (3), 183252.10.1016/0001-8686(83)80001-3CrossRefGoogle Scholar
Adamczyk, Z., Siwek, B., Weroński, P. & Jaszczółt, K. 2003 Particle deposition at electrostatically heterogeneous surfaces. Colloids Surf. A 222 (1–3), 1525.CrossRefGoogle Scholar
Anderson, D.M. & Davis, S.H. 1995 The spreading of volatile liquid droplets on heated surfaces. Phys. Fluids 7 (2), 248265.CrossRefGoogle Scholar
Anyfantakis, M. & Baigl, D. 2015 Manipulating the coffee-ring effect: interactions at work. ChemPhysChem 16 (13), 27262734.10.1002/cphc.201500410CrossRefGoogle ScholarPubMed
Anyfantakis, M., Geng, Z., Morel, M., Rudiuk, S. & Baigl, D. 2015 Modulation of the coffee-ring effect in particle/surfactant mixtures: the importance of particle–interface interactions. Langmuir 31 (14), 41134120.10.1021/acs.langmuir.5b00453CrossRefGoogle ScholarPubMed
Bhardwaj, R., Fang, X., Somasundaran, P. & Attinger, D. 2010 Self-assembly of colloidal particles from evaporating droplets: role of DLVO interactions and proposition of a phase diagram. Langmuir 26 (11), 78337842.10.1021/la9047227CrossRefGoogle ScholarPubMed
Bridonneau, N., Zhao, M., Battaglini, N., Mattana, G., Thévenet, V., Noël, V., Roché, M., Zrig, S. & Carn, F. 2020 Self-assembly of nanoparticles from evaporating sessile droplets: fresh look into the role of particle/substrate interaction. Langmuir 36 (39), 1141111421.10.1021/acs.langmuir.0c01546CrossRefGoogle ScholarPubMed
Brutin, D. & Starov, V. 2018 Recent advances in droplet wetting and evaporation. Chem. Soc. Rev. 47 (2), 558585.CrossRefGoogle ScholarPubMed
Coombs, N.C.J., Chubynsky, M.V. & Sprittles, J.E. 2024 a Colloidal deposits from evaporating sessile droplets: a computationally efficient framework for predicting the final deposit shape. Phys. Rev. E 110 (6), 064607.10.1103/PhysRevE.110.064607CrossRefGoogle ScholarPubMed
Coombs, N.C.J., Sprittles, J.E. & Chubynsky, M.V. 2024 b Colloidal deposits from evaporating sessile droplets: coffee ring versus surface capture. Phys. Rev. Fluids 9 (6), 064304.10.1103/PhysRevFluids.9.064304CrossRefGoogle Scholar
Copson, E.T. 1947 On the problem of the electrified disc. Proc. Edin. Math. Soc. 8 (1), 1419.10.1017/S0013091500027644CrossRefGoogle Scholar
D’Ambrosio, H.-M., Wilson, S.K., Wray, A.W. & Duffy, B.R. 2023 The effect of the spatial variation of the evaporative flux on the deposition from a thin sessile droplet. J. Fluid Mech. 970, A1.10.1017/jfm.2023.503CrossRefGoogle Scholar
D’Ambrosio, H.-M., Wray, A.W. & Wilson, S.K. 2025 a The effect of contact-line motion on the deposition of particles from an evaporating sessile droplet. J. Fluid Mech. (under review).Google Scholar
D’Ambrosio, H.-M., Wray, A.W. & Wilson, S.K. 2025 b The effect of particle–substrate adsorption on the deposition of particles from a thin evaporating sessile droplet. J. Engng Math. 151 (1), 1.10.1007/s10665-025-10423-0CrossRefGoogle Scholar
Deegan, R.D., Bakajin, O., Dupont, T.F., Huber, G., Nagel, S.R. & Witten, T.A. 1997 Capillary flow as the cause of ring stains from dried liquid drops. Nature 389 (6653), 827829.10.1038/39827CrossRefGoogle Scholar
Deegan, R.D., Bakajin, O., Dupont, T.F., Huber, G., Nagel, S.R. & Witten, T.A. 2000 Contact line deposits in an evaporating drop. Phys. Rev. E 62 (1), 756765.CrossRefGoogle Scholar
Devineau, S., Anyfantakis, M., Marichal, L., Kiger, L., Morel, M., Rudiuk, S. & Baigl, D. 2016 Protein adsorption and reorganization on nanoparticles probed by the coffee-ring effect: application to single point mutation detection. J. Am. Chem. Soc. 138 (36), 1162311632.CrossRefGoogle ScholarPubMed
Dugyala, V.R. & Basavaraj, M.G. 2014 Control over coffee-ring formation in evaporating liquid drops containing ellipsoids. Langmuir 30 (29), 86808686.10.1021/la500803hCrossRefGoogle ScholarPubMed
Eales, A.D., Routh, A.F., Dartnell, N. & Goddard, S. 2015 Evaporation of pinned droplets containing polymer – an examination of the important groups controlling final shape. AIChE J. 61 (5), 17591767.10.1002/aic.14777CrossRefGoogle Scholar
Elimelech, M. 1994 Particle deposition on ideal collectors from dilute flowing suspensions: mathematical formulation, numerical solution, and simulations. Sep. Tech. 4 (4), 186212.10.1016/0956-9618(94)80024-3CrossRefGoogle Scholar
Eral, H.B., Mampallil Augustine, D., Duits, M.H.G. & Mugele, F. 2011 Suppressing the coffee stain effect: how to control colloidal self-assembly in evaporating drops using electrowetting. Soft Matt. 7 (10), 49544958.10.1039/c1sm05183kCrossRefGoogle Scholar
Erdem, A.K., Denner, F. & Biancofiore, L. 2024 Numerical analysis of the dispersion and deposition of particles in evaporating sessile droplets. Langmuir 40 (26), 1342813445.10.1021/acs.langmuir.4c00680CrossRefGoogle ScholarPubMed
Gelderblom, H., Diddens, C. & Marin, A. 2022 Evaporation-driven liquid flow in sessile droplets. Soft Matt. 18 (45), 85358553.CrossRefGoogle ScholarPubMed
Golovko, D.S., Butt, H.-J. & Bonaccurso, E. 2009 Transition in the evaporation kinetics of water microdrops on hydrophilic surfaces. Langmuir 25 (1), 7578.10.1021/la803474xCrossRefGoogle ScholarPubMed
Guazzelli, É. & Pouliquen, O. 2018 Rheology of dense granular suspensions. J. Fluid Mech. 852, P1.10.1017/jfm.2018.548CrossRefGoogle Scholar
Harris, D.J., Hu, H., Conrad, J.C. & Lewis, J.A. 2007 Patterning colloidal films via evaporative lithography. Phys. Rev. Lett. 98 (14), 148301.10.1103/PhysRevLett.98.148301CrossRefGoogle ScholarPubMed
Hocking, L.M. 1983 The spreading of a thin drop by gravity and capillarity. Q. J. Mech. Appl. Math. 36 (1), 5569.10.1093/qjmam/36.1.55CrossRefGoogle Scholar
Howard, N.S., Archer, A.J., Sibley, D.N., Southee, D.J. & Wijayantha, K.G.U. 2023 Surfactant control of coffee ring formation in carbon nanotube suspensions. Langmuir 39 (3), 929941.10.1021/acs.langmuir.2c01691CrossRefGoogle ScholarPubMed
Hu, H. & Larson, R.G. 2002 Evaporation of a sessile droplet on a substrate. J. Phys. Chem. B 106 (6), 13341344.10.1021/jp0118322CrossRefGoogle Scholar
Issakhani, S., Jadidi, O., Farhadi, J. & Bazargan, V. 2023 Geometrically-controlled evaporation-driven deposition of conductive carbon nanotube patterns on inclined surfaces. Soft Matt. 19 (7), 13931406.10.1039/D2SM01431ACrossRefGoogle ScholarPubMed
Jiang, C., Zhong, Z., Liu, B., He, Z., Zou, J., Wang, L., Wang, J., Peng, J. & Cao, Y. 2016 Coffee-ring-free quantum dot thin film using inkjet printing from a mixed-solvent system on modified ZnO transport layer for light-emitting devices. ACS Appl. Mater. Interfaces 8 (39), 2616226168.10.1021/acsami.6b08679CrossRefGoogle ScholarPubMed
Kajiya, T. & M 2011 Dynamics of drying process of polymer solution droplets: analysis of polymer transport and control of film profiles. Nihon Reoroji Gakk. 39 (1_2), 1728.10.1678/rheology.39.17CrossRefGoogle Scholar
Kajiya, T., Kaneko, D. & M 2008 Dynamical visualization of coffee stain phenomenon in droplets of polymer solution via fluorescent microscopy. Langmuir 24, 1236912374.CrossRefGoogle ScholarPubMed
Kang, S.J., Vandadi, V., Felske, J.D. & Masoud, H. 2016 Alternative mechanism for coffee-ring deposition based on active role of free surface. Phys. Rev. E 94 (6), 063104.10.1103/PhysRevE.94.063104CrossRefGoogle Scholar
Kaplan, C.N. & Mahadevan, L. 2015 Evaporation-driven ring and film deposition from colloidal droplets. J. Fluid Mech. 781, R2.10.1017/jfm.2015.496CrossRefGoogle Scholar
Kim, J.Y., Gonçalves, M., Jung, N., Kim, H. & Weon, B.M. 2021 Evaporation and deposition of inclined colloidal droplets. Sci. Rep. 11, 17784.10.1038/s41598-021-97256-wCrossRefGoogle ScholarPubMed
Kumar, S., Basavaraj, M.G. & Satapathy, D.K. 2023 Effect of colloidal surface charge on desiccation cracks. Langmuir 39 (29), 1024910258.10.1021/acs.langmuir.3c01326CrossRefGoogle ScholarPubMed
Kurrat, R., Prenosil, J.E. & Ramsden, J.J. 1997 Kinetics of human and bovine serum albumin adsorption at silica–titania surfaces. J. Colloid Interface Sci. 185 (1), 18.10.1006/jcis.1996.4528CrossRefGoogle ScholarPubMed
Kurrat, R., Ramsden, J.J. & Prenosil, J.E. 1994 Kinetic model for serum albumin adsorption: experimental verification. J. Chem. Soc. Faraday Trans. 90 (4), 587590.10.1039/ft9949000587CrossRefGoogle Scholar
Larson, R.G. 2014 Transport and deposition patterns in drying sessile droplets. AIChE J. 60 (5), 15381571.10.1002/aic.14338CrossRefGoogle Scholar
Larsson, C. & Kumar, S. 2022 Quantitative analysis of the vertical-averaging approximation for evaporating thin liquid films. Phys. Rev. Fluids 7 (9), 094002.CrossRefGoogle Scholar
Li, H., Duan, Y., Shao, Z., Zhang, G., Li, H., Huang, Y. & Yin, Z. 2020 High-resolution pixelated light emitting diodes based on electrohydrodynamic printing and coffee-ring-free quantum dot film. Adv. Mater. Technol. 5 (10), 2000401.10.1002/admt.202000401CrossRefGoogle Scholar
Li, Y.-F., Sheng, Y.-J. & Tsao, H.-K. 2013 Evaporation stains: suppressing the coffee-ring effect by contact angle hysteresis. Langmuir 29 (25), 78027811.10.1021/la400948eCrossRefGoogle ScholarPubMed
Malinowski, R., Volpe, G., Parkin, I.P. & Volpe, G. 2018 Dynamic control of particle deposition in evaporating droplets by an external point source of vapor. J. Phys. Chem. Lett. 9 (3), 659664.10.1021/acs.jpclett.7b02831CrossRefGoogle Scholar
Malla, L.K., Bhardwaj, R. & Neild, A. 2019 Analysis of profile and morphology of colloidal deposits obtained from evaporating sessile droplets. Colloids Surf. A 567, 150160.10.1016/j.colsurfa.2019.01.028CrossRefGoogle Scholar
Mampallil, D. & Eral, H.B. 2018 A review on suppression and utilization of the coffee-ring effect. Adv. Colloid Interface Sci. 252, 3854.10.1016/j.cis.2017.12.008CrossRefGoogle ScholarPubMed
Mirbagheri, M. & Hwang, D.K. 2019 The coffee-ring effect on 3D patterns: a simple approach to creating complex hierarchical materials. Adv. Mater. Interfaces 6 (16), 1900003.10.1002/admi.201900003CrossRefGoogle Scholar
Molchanov, S.P., Roldughin, V.I., Chernova-Kharaeva, I.A. & Senchikhin, I.N. 2019 The effect of electrokinetic potential on evaporation of colloidal dispersion droplets. Colloid J. 81, 136145.10.1134/S1061933X19020091CrossRefGoogle Scholar
Moore, M.R., Vella, D. & Oliver, J.M. 2021 The nascent coffee ring: how solute diffusion counters advection. J. Fluid Mech. 920, A54.10.1017/jfm.2021.463CrossRefGoogle Scholar
Moore, M.R., Vella, D. & Oliver, J.M. 2022 The nascent coffee ring with arbitrary droplet contact set: an asymptotic analysis. J. Fluid Mech. 940, A38.10.1017/jfm.2022.251CrossRefGoogle Scholar
Moore, M.R. & Wray, A.W. 2023 Gravitational effects on coffee-ring formation during the evaporation of sessile droplets. J. Fluid Mech. 967, A26.10.1017/jfm.2023.493CrossRefGoogle Scholar
Murisic, N. & Kondic, L. 2011 On evaporation of sessile drops with moving contact lines. J. Fluid Mech. 679, 219246.10.1017/jfm.2011.133CrossRefGoogle Scholar
Nguyen, T.A.H., Hampton, M.A. & Nguyen, A.V. 2013 Evaporation of nanoparticle droplets on smooth hydrophobic surfaces: the inner coffee ring deposits. J. Phys. Chem. C 117 (9), 47074716.10.1021/jp3126939CrossRefGoogle Scholar
Oliver, J.M., Whiteley, J.P., Saxton, M.A., Vella, D., Zubkov, V.S. & King, J.R. 2015 On contact-line dynamics with mass transfer. Eur. J. Appl. Math. 26 (5), 671719.10.1017/S0956792515000364CrossRefGoogle Scholar
Orejon, D., Sefiane, K. & Shanahan, M.E.R. 2011 Stick–slip of evaporating droplets: substrate hydrophobicity and nanoparticle concentration. Langmuir 27 (21), 1283412843.10.1021/la2026736CrossRefGoogle ScholarPubMed
Ozawa, K., Nishitani, E. & M 2005 Modeling of the drying process of liquid droplet to form thin film. Jpn J. Appl. Phys. 44 (6R), 4229.10.1143/JJAP.44.4229CrossRefGoogle Scholar
Patil, N.D., Bange, P.G., Bhardwaj, R. & Sharma, A. 2016 Effects of substrate heating and wettability on evaporation dynamics and deposition patterns for a sessile water droplet containing colloidal particles. Langmuir 32 (45), 1195811972.10.1021/acs.langmuir.6b02769CrossRefGoogle ScholarPubMed
Popov, Y.O. 2005 Evaporative deposition patterns: spatial dimensions of the deposit. Phys. Rev. E 71 (3), 036313.10.1103/PhysRevE.71.036313CrossRefGoogle ScholarPubMed
Sáenz, P.J., Wray, A.W., Che, Z., Matar, O.K., Valluri, P., Kim, J. & Sefiane, K. 2017 Dynamics and universal scaling law in geometrically-controlled sessile drop evaporation. Nat. Commun. 8 (1), 14783.10.1038/ncomms14783CrossRefGoogle Scholar
Sefiane, K., Duursma, G. & Arif, A. 2021 Patterns from dried drops as a characterisation and healthcare diagnosis technique, potential and challenges: a review. Adv. Colloid Interface Sci. 298, 102546.CrossRefGoogle ScholarPubMed
Shao, X., Duan, F., Hou, Y. & Zhong, X. 2020 Role of surfactant in controlling the deposition pattern of a particle-laden droplet: fundamentals and strategies. Adv. Colloid Interface Sci. 275, 102049.10.1016/j.cis.2019.102049CrossRefGoogle ScholarPubMed
Siregar, D.P. 2012 Numerical simulation of evaporation and absorption of inkjet printed droplets. PhD thesis, Technische Universiteit Eindhoven.Google Scholar
Siregar, D.P., Kuerten, J.G.M. & Van der Geld, C.W.M. 2013 Numerical simulation of the drying of inkjet-printed droplets. J. Colloid Interface Sci. 392, 388395.10.1016/j.jcis.2012.09.063CrossRefGoogle ScholarPubMed
Smalyukh, I.I., Zribi, O.V., Butler, J.C., Lavrentovich, O.D. & Wong, G.C.L. 2006 Structure and dynamics of liquid crystalline pattern formation in drying droplets of DNA. Phys. Rev. Lett. 96 (17), 177801.10.1103/PhysRevLett.96.177801CrossRefGoogle ScholarPubMed
Smith, F.R. & Brutin, D. 2018 Wetting and spreading of human blood: recent advances and applications. Curr. Opin. Colloid Interface Sci. 36, 7883.10.1016/j.cocis.2018.01.013CrossRefGoogle Scholar
Spielman, L.A. & Friedlander, S.K. 1974 Role of the electrical double layer in particle deposition by convective diffusion. J. Colloid Interface Sci. 46 (1), 2231.CrossRefGoogle Scholar
Sung, P.-F., Wang, L. & Harris, M.T. 2019 Deposition of colloidal particles during the evaporation of sessile drops: dilute colloidal dispersions. Intl J. Chem. Engng 2019, 7954965.Google Scholar
Weber, H. 1873 Über die Besselschen Functionen und ihre Anwendung auf die Theorie der elektrischen ströme. J. für Die Reine Angew. Math. 75, 75105.Google Scholar
Weon, B.M. & Je, J.H. 2013 Self-pinning by colloids confined at a contact line. Phys. Rev. Lett. 110 (2), 028303.10.1103/PhysRevLett.110.028303CrossRefGoogle Scholar
Widjaja, E. & Harris, M.T. 2008 Particle deposition study during sessile drop evaporation. AIChE J. 54 (9), 22502260.10.1002/aic.11558CrossRefGoogle Scholar
Wilson, S.K. & D’Ambrosio, H.-M. 2023 Evaporation of sessile droplets. Annu. Rev. Fluid Mech. 55, 481509.10.1146/annurev-fluid-031822-013213CrossRefGoogle Scholar
Wray, A.W. & Moore, M.R. 2023 Evaporation of non-circular droplets. J. Fluid Mech. 961, A11.10.1017/jfm.2023.229CrossRefGoogle Scholar
Wray, A.W. & Moore, M.R. 2024 High-order asymptotic methods provide accurate, analytic solutions to intractable potential problems. Sci. Rep. 14 (1), 4225.10.1038/s41598-024-54377-2CrossRefGoogle ScholarPubMed
Wray, A.W., Wray, P.S., Duffy, B.R. & Wilson, S.K. 2021 Contact-line deposits from multiple evaporating droplets. Phys. Rev. Fluids 6 (7), 073604.CrossRefGoogle Scholar
Yan, Q., Gao, L., Sharma, V., Chiang, Y.-M. & Wong, C.C. 2008 Particle and substrate charge effects on colloidal self-assembly in a sessile drop. Langmuir 24 (20), 1151811522.10.1021/la802159tCrossRefGoogle Scholar
Yunker, P.J., Still, T., Lohr, M.A. & Yodh, A.G. 2011 Suppression of the coffee-ring effect by shape-dependent capillary interactions. Nature 476 (7360), 308311.10.1038/nature10344CrossRefGoogle ScholarPubMed
Zhang, J., Sibley, D.N., Tseluiko, D. & Archer, A.J. 2024 Dynamics of particle aggregation in dewetting films of complex liquids. J. Fluid Mech. 990, A10.10.1017/jfm.2024.466CrossRefGoogle Scholar
Zhou, P., Yu, H., Zou, W., Wang, Z. & Liu, L. 2019 High-resolution and controllable nanodeposition pattern of ag nanoparticles by electrohydrodynamic jet printing combined with coffee ring effect. Adv. Mater. Interfaces 6 (20), 1900912.10.1002/admi.201900912CrossRefGoogle Scholar
Zigelman, A. & Manor, O. 2018 a The deposition of colloidal particles from a sessile drop of a volatile suspension subject to particle adsorption and coagulation. J. Colloid Interface Sci. 509, 195208.10.1016/j.jcis.2017.08.088CrossRefGoogle ScholarPubMed
Zigelman, A. & Manor, O. 2018 b Simulations of the dynamic deposition of colloidal particles from a volatile sessile drop. J. Colloid Interfae Sci. 525, 282290.10.1016/j.jcis.2018.04.054CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. A droplet of volatile fluid lying on a solid substrate evaporates into the surrounding gas. The droplet contains inert particles that are transported by the competing effects of advection (yellow), diffusion (blue) and substrate adsorption (magenta). Advection to the pinned contact line is driven by the flow due to the evaporation. This increases the local particle volume fraction, driving a diffusive flux of particles near the contact line. Finally, particles may be arrested on the substrate due to adsorption.

Figure 1

Figure 2. Accumulated mass flux into the contact line $\mathcal{M}(\theta )$ for $\mathcal{V} = 0.25, 1, 4, 8$ (greyscales) and the equivalent coefficient in the no-substrate adsorption problem $\mathcal{M}_{\textit{MVO}}(\theta )$ (blue). The dashed red curves represent the small-time and large-time limits (3.9) and (3.10).

Figure 2

Figure 3. The coefficient $A(\theta )$ for $\mathcal{V} = 0.25, 1, 4, 8$ (greyscales) and the equivalent coefficient in the no-substrate adsorption problem $A_{\textit{MVO}}(\theta )$ (blue). The dashed red curves represent the small-time and large-time limits (3.26) and (3.27).

Figure 3

Figure 4. The leading-order composite suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 219$, ${\textit{Da}} = 154$ ($\mathcal{V} \approx 1.1$). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. The red dashed curves indicate the corresponding numerical solution of (2.33)–(2.37), while the arrows indicate increasing evaporation time.

Figure 4

Figure 5. The cumulative mass distribution in the suspension (orange-scale curves) and deposited onto the substrate (blue-scale curves) for (a) ${\textit{Pe}} = 219,\ {\textit{Da}} = 154$ ($\mathcal{V} \approx 1.1$) and (b) ${\textit{Pe}} = 40,\ {\textit{Da}} = 60$ ($\mathcal{V} \approx 2.4$). Results are displayed at $30\,\%,\ 60\,\%$ and $90\,\%$ of the drying time.

Figure 5

Figure 6. Percentages of the total mass in suspension (black) and deposited (red) in the droplet bulk and in suspension (blue) and deposited (magenta) near the contact line as a function of drying time. In each case, ${\textit{Pe}} = 40$ and we take ${\textit{Da}} =$ (a) $10$ ($\mathcal{V} \approx 0.39$), (b) $20$ ($\mathcal{V} \approx 0.79$), (c) $60$ ($\mathcal{V} \approx 2.36$) and (d) $100$ ($\mathcal{V} \approx 3.92$).

Figure 6

Figure 7. The adsorbed mass per unit area in the droplet bulk $\mathcal{D}_{{b}}$ (solid curves) and contact-line region $\mathcal{D}_{{cl}}$ (dashed curves) for ${\textit{Pe}} = 20$ (black), $40, 60, 80, 100$ (light grey).

Figure 7

Figure 8. The leading-order suspension (left) and deposited (right) mass in the inner region as a function of $R = {\textit{Pe}}^2(1-r)$ for $(a)$$\mathcal{V}=0.25$, $(b)$$\mathcal{V}=1$, $(c)$$\mathcal{V}=4$ and $(d)$$\mathcal{V}=8$. In each case, we show profiles at $25\,\%$$(\theta = 0.75\theta _0)$, $50\,\%$$(\theta = 0.5\theta _0)$ and $75\,\%$$(\theta = 0.25\theta _0)$ of drying time: drying time is indicated by the arrow in each figure.

Figure 8

Figure 9. The onset of jamming (bluescale curves, left-hand axis) and deposited layer effects (redscale curves, right-hand axis) as a percentage of the total drying time given by the roots of (3.41) and (3.46), respectively. For each case, we vary the Péclet number, ${\textit{Pe}} = 5,\ 10,\ 20,\ 40$, while for $\theta _{{ct}}$, we take $R = 1$ as an illustrative value.

Figure 9

Figure 10. The leading-order composite suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 40$, ${\textit{Da}} = 5$ ($\mathcal{V} \approx 0.2$). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. The red dashed curves indicate the corresponding numerical solution of (2.33)–(2.37), while the arrows indicate increasing evaporation time.

Figure 10

Figure 11. Numerical results for the suspension (a,b) and deposited (c,d) mass as a function of time for data taken from Kurrat et al. (1997) and Hu & Larson (2002), for which ${\textit{Pe}} = 6.72$, ${\textit{Da}} = 1.03$ ($\mathcal{V} \approx 0.24$). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. Arrows indicate increasing evaporation time.

Figure 11

Figure 12. Numerical results for the suspension (a,b) and deposited (c,d) mass as a function of time for ${\textit{Pe}} = 1$, ${\textit{Da}} = 0.5$ ($\mathcal{V} \approx 0.79$). The initial profile mirrors the droplet shape and is shown as the bold blue line, while profiles at $15\,\%$ intervals of the drying time are shown in increasingly lighter shades of grey. A close-up of the nascent coffee-ring profile is shown in (b,d), whereby we see the characteristic formation of the peak just inside the contact line. Arrows indicate increasing evaporation time.

Figure 12

Figure 13. A comparison between simulation data from figure 9(a) in Widjaja & Harris (2008) (circles), the leading-order composite asymptotic adsorbed mass given by (3.33) (solid grey curve) and the numerical solution of (2.33)–(2.37) (dashed red curve). In the language of the present paper, the results are presented for ${\textit{Da}} = {\textit{Pe}} = 12.5$$(\mathcal{V} = \pi /2)$.

Figure 13

Figure 14. Comparison between the leading-order deposited mass in the contact-line region at the time of contact-line de-pinning $\mathcal{H}_{{cl}}(\theta ^*)$ given by (3.36) for (a) ${\textit{Pe}}\approx 50.0$, ${\textit{Da}}\approx 3.66$ and $\mathcal{V}\approx 0.115$ (black line) and ${\textit{Pe}}\approx 69.1$, ${\textit{Da}}\approx 4.47$ and $\mathcal{V}\approx 0.102$ (grey line) and the experimental data for the final mass in the ring from supplementary figures S8 and S9 in Bridonneau et al. (2020) for untreated (star) and aluminium-coated (diamond) particles, and for (b) ${\textit{Pe}}\approx 127$, ${\textit{Da}}\approx 20.5$ and $\mathcal{V}\approx 0.253$ (black line) and ${\textit{Pe}}\approx 213$, ${\textit{Da}}\approx 8.23$ and $\mathcal{V}\approx 0.061$ (grey line) and experimental data for the final mass in the ring from figures 6(a) and 6(b) in Bhardwaj et al. (2010) for $\text{pH}=11.7$ (circle) and $\text{pH}=2.8$ (square). The dashed lines correspond to the predicted mass in the contact-line region when the droplet evaporates with a pinned contact line throughout evaporation $\mathcal{H}_{{cl}}(0)$.