We develop a model for the thermodynamics and evaporation dynamics of aerosol droplets of a liquid, such as water, surrounded by gas. When the temperature and the chemical potential (or equivalently the humidity) are such that the vapor phase is in the thermodynamic equilibrium state, then, of course, droplets of the pure liquid evaporate over a relatively short time. However, if the droplets also contain nanoparticles or any other non-volatile solute, then the droplets can become thermodynamically stable. We show that the equilibrium droplet size depends strongly on the amount and solubility of the nanoparticles within, i.e., on the nature of the particle interactions with the liquid and, of course, also on the vapor temperature and chemical potential. We develop a simple thermodynamic model for such droplets and compare predictions with results from a lattice density functional theory that takes as input the same particle interaction properties, finding very good agreement. We also use dynamical density functional theory to study the evaporation/condensation dynamics of liquid from/to droplets as they equilibrate with the vapor, thereby demonstrating droplet stability.

An aerosol droplet is a small liquid drop in the colloidal size range of tens of nanometers up to a few micrometers in diameter, suspended in a gas like air. The lifetime of such an aerosol droplet is determined by a competition between gravity and evaporation.1,2 Larger droplets sediment from a height of the order of 2 m in less than a few seconds, while smaller droplets evaporate rapidly, as long as the temperature and pressure conditions are such that the vapor of the volatile liquid is in the thermodynamic equilibrium phase. In the case of water, such small droplets evaporate completely in the order of seconds. Wells, who was interested in the spreading of diseases through aerosols,1 showed that the average sedimentation time scales with 1/R02, where R0 is the initial droplet radius, while the evaporation time scales with R02. Therefore, a droplet of pure water always either sediments or evaporates sufficiently quickly, preventing it from traveling a significant horizontal distance from the source. In contrast to the droplets described by these simple estimates, aerosols produced by people when they breathe, talk, cough, or sneeze can stay airborne and, therefore, are potentially dangerous for much longer periods, of the order of hours instead of seconds.

This is because if the droplets contain nanoparticles, e.g., pathogenic germs, suspended within the liquid or some other non-volatile solute, then a sufficient quantity of these can stabilize the droplets for hours or even indefinitely.3,4 In other words, once such a droplet is formed, there may be some fast initial evaporation of the solvent liquid, but in the long time limit, a finite-size droplet remains that can stay airborne. The droplet is stabilized by the sufficiently large concentration of nanoparticles within. In the equilibrium state, the rate at which liquid evaporates from this droplet is balanced by the condensation of the vapor phase onto the droplet. Clearly, the temperature and relative humidity of the vapor atmosphere have to be taken into account.3,4

Therefore, studying such droplets allows us to address questions such as: How long does a droplet of saliva survive in the air before it evaporates? How does this depend on the amount of solute within? These questions are particularly relevant if any of that solute is an infectious virus, such as COVID or influenza. A related additional question of particular interest here is: What influence do the interactions between the non-volatile material within and the surrounding liquid have on the droplet size? Saliva droplets contain a significant amount of material other than water, including various mucus components, proteins, salts, and sometimes virus particles.4–6 The fact that this non-volatile material can stabilize the droplets goes some way toward explaining why it was observed that influenza viruses remain stable and infectious in aerosols across a wide range of relative humidities.7 Note that the air humidity is directly related to the vapor pressure or, equivalently, is determined by the chemical potential of the vapor. Because these quantities are so closely connected, we refer to them here almost interchangeably.

Aerosol droplets are not just relevant to the spread of disease; they are ubiquitous in the global environment, playing a crucial role in atmospheric science and meteorology. For example, aerosol droplets play a role in determining how long clouds persist in the sky before the water they contain returns to the ground as rain.8 They are also often the host locations for chemical reactions between airborne species,9 and the lifetime and stability of aerosol droplets must be considered when addressing the question of how long harmful chemicals and other pollutants remain in the atmosphere.10 

In what follows, we refer to all of the non-volatile material that may be found in an aerosol liquid droplet collectively as “nanoparticles.” Droplets of pure water evaporate in the air, but droplets containing sufficient numbers of nanoparticles are stabilized by the particles within. The recent study by Netz3 addressed this and many of the related issues to do with the evaporation of saliva droplets in air. When considering the thermodynamic stability of saliva droplets, Netz used a simple thermodynamics for mixtures, based on assuming ideal-mixing and ideal-volume additivity. Here, we go beyond this to develop a model to determine the equilibrium size of such droplets and how the droplet size depends on the humidity and temperature of the surrounding gas and also on the nature of the interactions of the solute particles with the solvent liquid. We show that nanoparticles that have a higher solubility (i.e., have a greater preference for being dispersed in the liquid) lead to larger droplets. We develop a simple capillarity-approximation based thermodynamic model that we validate by comparing with the results from a lattice density functional theory (DFT), which is a theory for the microscopic density profile of the liquid and nanoparticles within the droplet.11,12 We also study the evaporation/condensation dynamics of the liquid from/onto non-equilibrium droplets using dynamical density functional theory (DDFT),12–15 thereby demonstrating the stability of the droplets and elucidating the properties of the formation dynamics. By using DDFT, we assume that the motion of all particles in the system is diffusive, stemming from our expectation that the droplet dynamics is dominated by the diffusive exchange of solvent molecules from the liquid to the vapor phase and also that all motion within the droplet can be treated as diffusive. In our work here, we treat the surrounding gas as being solely made up of the vapor phase and do not explicitly treat the inert gas molecules that are present, e.g., in the atmosphere. However, since these largely play the role of passive spectators, the results presented here also apply to the case of aerosol water droplets in air.

The lattice DFT that we use here and variants of it have previously been successfully applied to the study of various properties of liquids condensing and adsorbed in pores and porous media16–18 and to liquid droplets on surfaces.19–22 The recent study in Ref. 23 is particularly noteworthy because comparisons with experimental measurements on water droplets containing nanoparticles drying on various surfaces were made, demonstrating good agreement between the theory and the experiments. In view of this, we are confident that the lattice DFT and DDFT used here also provide a good description of the thermodynamics and dynamics of aerosol droplets. Additionally, in a future study, we will support our lattice DFT results with additional off-lattice DFT calculations.

This paper proceeds as follows: In Sec. II, we develop a generalized lattice-gas model for nanoparticle laden droplets and present the lattice DFT that we use to determine the density profiles of the liquid and nanoparticles within the droplets. This theory also yields all the relevant thermodynamic quantities pertaining to the droplets. Bulk thermodynamic quantities are also briefly discussed in this section. In Sec. III, we present results for the liquid and nanoparticle density profiles in droplets calculated using our DFT model. In Sec. IV, we then present our capillarity-approximation based thermodynamic model, comparing it with DFT results for the total amount of liquid in the droplets, as quantities like the number of nanoparticles, humidity, and interaction strength between the nanoparticles and the liquid are varied. In Sec. V, we present our DDFT model and results for the formation dynamics of droplets. Then, in Sec. VI, we compare results for droplet dynamics obtained from DDFT with the dynamics from Picard iteration, which is a fictitious dynamics used to solve the equations of DFT. This section will be of interest to DFT practitioners but may be skipped by those from a more general audience. Finally, in Sec. VII, we make a few concluding remarks.

In Fig. 1, we illustrate a cross-section through an aerosol liquid droplet containing nanoparticles. To model this, we discretize the system onto a lattice, as also illustrated in Fig. 1. We choose the lattice spacing to be the diameter of the nanoparticles, σ. Therefore, lattice sites that contain a nanoparticle have just one of them within, while lattice sites that are “full” of the liquid have a large number of liquid molecules within. As an example of a concrete physical system consisting of the case where the solvent is water and for σ ≈ 100 nm (roughly the diameter of a COVID virion), we would find 3.3×107 molecules in each lattice site “full” of the water. Following Ref. 21, this discretization onto a lattice allows us to map the system onto a two-species lattice-gas (generalized Ising model) and thereby to investigate the thermodynamics of the system. The resulting model can also be thought of as a discretized partial differential equation for the density distributions of the liquid and nanoparticles.24 

FIG. 1.

On the left is an illustration of the system of interest, namely an aerosol liquid droplet containing nanoparticles that is surrounded by gas (i.e., the vapor phase). The red circles represent the suspended nanoparticles. To treat this, we coarse-grain the system onto a square lattice, as illustrated on the right. We choose the size of each lattice site to correspond roughly to the diameter of the nanoparticles. Lattice sites are described as either occupied with a nanoparticle (red circle), occupied by liquid (blue square), or empty (white). We also define effective interaction potentials between pairs of lattice sites that represent coarse-grained analogs of the inter-particle interaction potentials in the original system on the left. Therefore, the liquid within the drop has the majority of lattice sites full of either liquid or nanoparticles, while the vapor outside has most (but not all) sites empty.

FIG. 1.

On the left is an illustration of the system of interest, namely an aerosol liquid droplet containing nanoparticles that is surrounded by gas (i.e., the vapor phase). The red circles represent the suspended nanoparticles. To treat this, we coarse-grain the system onto a square lattice, as illustrated on the right. We choose the size of each lattice site to correspond roughly to the diameter of the nanoparticles. Lattice sites are described as either occupied with a nanoparticle (red circle), occupied by liquid (blue square), or empty (white). We also define effective interaction potentials between pairs of lattice sites that represent coarse-grained analogs of the inter-particle interaction potentials in the original system on the left. Therefore, the liquid within the drop has the majority of lattice sites full of either liquid or nanoparticles, while the vapor outside has most (but not all) sites empty.

Close modal
We denote the location of each lattice site by the index i. In three dimensions, we have i = (i, j, k), where i, j, and k are integers. To simplify our calculations below, we assume instead that our system is two-dimensional (2D) and, therefore, we have i = (i, j). We introduce two occupation numbers for each lattice site, pin and pil, for the nanoparticles and for the liquid, respectively. If a lattice site is empty, then both pin=pil=0. If lattice site i is occupied by a nanoparticle, then pin=1, while pil=0. If instead lattice site i is full of liquid, then pin=0 and pil=1. We assume that it is impossible for both occupation numbers to simultaneously equal one. The potential energy of the system can then be approximated as
(1)
where the first three terms arise from the interactions between pairs of occupied lattice sites, while the last two terms are the contributions from any external potentials Φil and Φin on the liquid and nanoparticles, respectively. Here, we assume Φil=Φin=0. Note that i=i=1Mxj=1My, i.e., this denotes a sum over all lattice sites in the system, where Mx is the number of lattice sites along the Cartesian x-direction indexed by i and My is the number along the y-direction, indexed by j. Similarly, ∑i,j denotes a sum over all pairs of lattice sites. The pair interaction terms involved the discretized pair-potentials εijll, εijnl and εijnn, which we assume are given by the matrices εijll=εllcij, εijnl=εnlcij, and εijnn=εnncij, where the parameters ɛll, ɛnl, and ɛnn determine the overall strength of the pair-interactions and the matrix
(2)
where NNi denotes the lattice sites that are nearest neighbors to site i, while NNNi indicates the sites that are the next nearest neighbors. These three short-ranged (truncated) potentials are chosen to have the above form for the reasons discussed in Refs. 21 and 24, namely that this choice greatly reduces the influence of the approximation of having discretized the system onto a lattice. If other values for the entries of the matrix cij were used, we would obtain, e.g., non-circular droplets and other similar influences from the underlying grid. There is a corresponding choice one can make when applying the model in 3D.21,22,25 Note that with the sign convention used in Eq. (1) for the pair interaction terms, positive values of ɛll, ɛnl, and ɛnn correspond to attractive interactions between neighboring particles, while negative values (not considered here) correspond to repulsive interactions. The parameter ɛnl determines the overall strength of attraction between a nanoparticle and any liquid surrounding it. Therefore, varying this parameter most directly determines the free energy to insert a nanoparticle into the liquid,26–28 which is the quantity that determines the solubility of the nanoparticles in the liquid.
Having defined the Hamiltonian (1), one could proceed, e.g., by performing Monte-Carlo computer simulations, as was performed in Ref. 25. However, here we prefer to follow, e.g., Refs. 19–21, to develop a theory for the ensemble-averaged densities,
(3)
Therefore, we apply an extension of DFT11,12 to lattice-systems. We use the following approximation for the Helmholtz free energy of the system:21,29
(4)
where kB is Boltzmann’s constant and T is the temperature. To determine the equilibrium density profiles ρil and ρin corresponding to an aerosol droplet, we minimize the free energy F in the semi-grand ensemble, i.e., with a fixed total number of nanoparticles in the system
(5)
but with the chemical potential of the liquid μμl (or equivalently the relative humidity) as an input so that the total amount of liquid in the system
(6)
is an output of our calculations. We calculate the equilibrium density profiles using a Picard iteration scheme similar to that described in Ref. 19. Since we are in the semi-grand ensemble, the equilibrium density profiles {ρil,ρin} are those that minimize the semi-grand free energy,
(7)
subject to the constraint that Nn, given by (5), is equal to the desired value, N̂n. In other words, we solve the coupled set of equations
(8)
(9)
(10)
Differentiating Ω with respect to ρil and ρin and rearranging gives the conditions
(11)
(12)
(13)
We solve this set of equations iteratively, starting from an initial guess for the profiles. A standard Picard solver would take the current state ρim,old and replace it with the result of evaluating the right hand sides of Eqs. (11) and (12) with these densities, denoted ρim,rhs. Here and below, m ∈ {l, n}. Since we are working in a semi-grand ensemble where Nn is fixed, we enforce this by renormalizing the density profile of the nanoparticles {ρin} at each step so that Eq. (13) is satisfied. Note also that it is often necessary to mix the results from the previous step ρim,old with the results from evaluating the right hand sides, ρim,rhs,
(14)
where α is typically small, e.g., 0.01 ≤ α ≤ 0.1. This mixing increases the robustness of the scheme, in particular by preventing ρim,new from lying outside of the range (0,1).
Before presenting results from this lattice-DFT model, we briefly discuss a few relevant aspects of the bulk thermodynamics and phase behavior of the system. From Eq. (4), we find that for a uniform system with ρil=ρl=Nl/V and ρin=ρn=Nn/V, constants for all i, the Helmholtz free energy per unit volume, f = F/V, where V is the volume of the system, is given by
(15)
From this, we can obtain the following relation between the chemical potential of the liquid and the particle densities:
(16)
The pressure (or equivalently minus the semi-grand potential density) is obtained from the relation
(17)
For the case where no nanoparticles are present, i.e., where ρn → 0, the system exhibits vapor–liquid phase separation. The critical point occurs for βɛll = 2/3, where β=(kBT)1; i.e., the critical point is at (ρl,T)=(ρcl,Tc)=(12,3εll2kB). Bulk liquid–vapor phase coexistence occurs for μ = −3ɛll. Adding nanoparticles to the system can generally completely change the bulk phase behavior (see, e.g., Ref. 30). However, for the interaction parameter values considered here, in general, the effect of the nanoparticles is to somewhat shift the coexisting density and chemical potential values, but the overall qualitative phase behavior remains the same.

As mentioned earlier, to simplify our DFT calculations, we treat the system as being 2D, but our results could fairly easily be repeated in 3D at the expense of more computational cost. We consider the system with βɛll = 1.2, where the densities of the coexisting (pure) liquid and gas are ρl = 0.034 and ρl = 0.966. For the interaction strengths with the nanoparticles, we initially set βɛnn = 0.9 and βɛnl = 1.5. This choice has ɛnn < ɛll < ɛnl, which ensures that the nanoparticles prefer to stay well-dispersed within the liquid and do not aggregate. If ɛnn were larger and/or if ɛnl were smaller, this could potentially lead to the nanoparticles and the liquid demixing.30 In Sec. IV below, we present results for different values of ɛnl, which is the parameter that most directly determines the solubility of the nanoparticles in the liquid.

In Fig. 2, we present both the liquid and nanoparticle density profiles for the case when the total number of nanoparticles in the system is set to be Nn = 300 in a square box of size 55σ × 55σ. In the top row are displayed results for the case when βμ = −3.8, in the middle row for βμ = −4.5, and in the bottom row for βμ = −5. For the temperature considered here (i.e., for the value of βɛll = 1.2, used here), the chemical potential at gas–liquid coexistence is βμcoex = −3.6, and so these three chemical potential values correspond to relative humidity values of Hr = 80%, Hr = 38%, and Hr = 23%, respectively. If this were for the case of water in the air, then these Hr values correspond roughly speaking to a moist environment, a typical comfortable day-to-day level, and to a fairly dry level, where one may need extra moisturizing cream on skin. Note that the relative humidity is defined as Hr(μ) = 100 × p(μ)/p(μcoex), where the corresponding pressures are calculated via Eqs. (15)(17). We see that when the chemical potential is closer to the value at phase-coexistence (more humid air), the equilibrium droplet size is much larger than in dry air, which of course is to be expected.

FIG. 2.

Top row: the density profiles (liquid on the left, nanoparticles on the right) for an equilibrium nanoparticle laden droplet. The system interaction parameters are βɛll = 1.2, βɛnn = 0.9, βɛnl = 1.5, and liquid chemical potential βμ = −3.8 (corresponding to moist air with Hr = 80%), while the total number of nanoparticles in the system is constrained to be Nn = 300. Middle row: the corresponding case for βμ = −4.5 (comfortable moisture level, with Hr = 38%), and all other parameters the same. Bottom row: the corresponding case for βμ = −5 (dry air, with Hr = 23%), again with all other parameters the same. Note that the droplet decreases in size with decreasing μ, i.e., with decreasing air humidity. Moreover, the liquid density within the droplet is also lower.

FIG. 2.

Top row: the density profiles (liquid on the left, nanoparticles on the right) for an equilibrium nanoparticle laden droplet. The system interaction parameters are βɛll = 1.2, βɛnn = 0.9, βɛnl = 1.5, and liquid chemical potential βμ = −3.8 (corresponding to moist air with Hr = 80%), while the total number of nanoparticles in the system is constrained to be Nn = 300. Middle row: the corresponding case for βμ = −4.5 (comfortable moisture level, with Hr = 38%), and all other parameters the same. Bottom row: the corresponding case for βμ = −5 (dry air, with Hr = 23%), again with all other parameters the same. Note that the droplet decreases in size with decreasing μ, i.e., with decreasing air humidity. Moreover, the liquid density within the droplet is also lower.

Close modal

In Fig. 3, we display results for a case where ɛnl is set to be much lower than for the cases in Fig. 2. This corresponds to the nanoparticles having a fairly poor solubility in the liquid. As a consequence, we see from the density profiles that the nanoparticles gather to form a dense clump with only a relatively small amount of liquid within. Interestingly, the density of the liquid is highest around the edge of the droplet. We present this particular result to illustrate that the density profiles are not always as simple as those presented in Fig. 2. Before discussing any further DFT results and in particular how the droplet size varies depending on the number of nanoparticles within and the value of the parameter ɛnl, we first present our capillarity-approximation based thermodynamic model in Sec. IV below. Our aim is to compare results from the microscopic DFT model with this mesoscopic capillarity-approximation model. We show in the following section that the two are in excellent agreement.

FIG. 3.

Density profiles (liquid on the left, nanoparticles on the right), for the case when βɛll = 1.2, βɛnn = 0.9, βɛnl = 0.6 and liquid chemical potential βμ = −4 (corresponding to Hr = 65%), while the total number of nanoparticles in the system is constrained to be Nn = 1000. The smaller value of ɛnl used here (compared to that used in Fig. 2) results in a high nanoparticle density within the drop, and interestingly, the liquid density profile is highest on the edge of the droplet.

FIG. 3.

Density profiles (liquid on the left, nanoparticles on the right), for the case when βɛll = 1.2, βɛnn = 0.9, βɛnl = 0.6 and liquid chemical potential βμ = −4 (corresponding to Hr = 65%), while the total number of nanoparticles in the system is constrained to be Nn = 1000. The smaller value of ɛnl used here (compared to that used in Fig. 2) results in a high nanoparticle density within the drop, and interestingly, the liquid density profile is highest on the edge of the droplet.

Close modal

Since our DFT calculations in the previous section were for a 2D system, here we present our capillarity approximation (CA) model in 2D. However, it is no more difficult to construct it in 3D, so at various points in the following, we additionally give the corresponding 3D equations.

Consider first a system of volume V (strictly, area in 2D) containing just the bulk vapor phase. The grand potential of the system is then just
(18)
where pvap is the bulk pressure in the vapor phase. In our model, we assume this is given by Eq. (17) in the limit ρn → 0. Note that the vapor density ρl is determined by the chemical potential μ, which we assume to be specified. Alternatively, one could assume that the vapor density (i.e., the humidity) is given, and then the chemical potential can be calculated via Eq. (16) with ρn → 0.
Now consider the case when the system also contains a circular (in 2D) droplet of radius R, surrounded by the vapor phase; see, e.g., Fig. 1. The grand potential can now be approximated as
(19)
The first term is the bulk contribution due to the droplet, which has volume (area in 2D) equal to πR2; cf. Eq. (18). The second term is the corresponding contribution due to the vapor filling the remainder of the system. The final term is the contribution from the interface between the liquid and gas phases; γ is the interfacial tension. The quantity of relevance in the calculations that follow is the difference between these, ΔΩ = Ωdrop − Ωvap, which is given by
(20)
In 3D, the above should be replaced by the following:
(21)
Here we assume that the interfacial tension γ is that of the pure vapor–liquid system at the given temperature. For our model, when βɛll = 1.2, we obtain βγσ = 0.685 87, which is calculated in the usual way (see, e.g., Refs. 11 and 19). The fixed number of nanoparticles in the droplet means that the nanoparticle density in the droplet is (in 2D)
(22)
or (in 3D)
(23)
Substituting Eq. (22) into Eq. (17) [together with Eq. (15)], we obtain the following expression for the pressure in the drop (in 2D):
(24)
Substituting this into Eq. (20), we obtain an expression for ΔΩ = ΔΩ(ρl, R), which is a function of (ρl, R). The equilibrium droplet radius R is the value that minimizes the free energy; therefore, we require
(25)
Similarly, substituting Eq. (22) into Eq. (16) yields the following second equation:
(26)
that is also a function of the two unknowns (ρl, R). This equation corresponds to requiring that the liquid density in the drop equal the higher of the two possible values determined by the selected value of the chemical potential μ. In other words, we require the value of the chemical potential within the droplet to be the same as that in the surrounding vapor. We then simply solve numerically (using fsolve in Maple) the pair of simultaneous Eqs. (25) and (26) for the two unknowns, ρl and R. From these, we can then easily obtain the total amount of liquid in the droplet as (in 2D)
(27)
or (in 3D)
(28)
In Fig. 4, we compare the results from this simple CA with results from the DFT for the case when βɛll = 1.2, βɛnn = 0.9, βɛnl = 1.5 and for various values of the chemical potential μ (i.e., for various values of the humidity). To compare, we plot Γl from the CA in Eq. (27) together with the result from the DFT, where we determine the amount of liquid in the droplet via the following sum over lattice sites within the droplet,
(29)
where ρmaxl=sup({ρil}) is the maximum liquid density in the system (i.e., in the center of the drop). Therefore, we do not include the material in the surrounding vapor when determining the mass of liquid in the droplet. Choosing the threshold as ρil>0.99ρmaxl is somewhat arbitrary. However, changing this to, e.g., ρil>0.9ρmaxl only slightly changes our results. From the results in Fig. 4, we see that the agreement between the CA model and the DFT is rather good. We believe the main source of error in our CA model is in our decision to use the value for the interfacial tension γ obtained for the pure liquid system. In reality, the presence of the nanoparticles changes the value of γ. However, given the agreement we see in Fig. 4, we conclude that, at least for the present system, this is a reasonable approximation to make.
FIG. 4.

Plots of the amount of liquid Γl in the droplet as a function of the number of nanoparticles in the droplet Nn for various values of the chemical potential μ, i.e., for varying humidity. We compare results from the capillarity approximation (CA), where Γl is calculated via Eq. (27), with those from DFT, where Γl is calculated via Eq. (29). The results here are for βɛll = 1.2, βɛnn = 0.9, and βɛnl = 1.5. Note that βμ = −3.8 corresponds to a relative humidity of Hr = 80%; βμ = −4 corresponds to Hr = 65%; βμ = −5 corresponds to Hr = 23%; and βμ = −6 corresponds to Hr = 8%. Note also that the CA lines for βμ = −5 and βμ = −6 are barely visible, since the corresponding DFT result is almost identical.

FIG. 4.

Plots of the amount of liquid Γl in the droplet as a function of the number of nanoparticles in the droplet Nn for various values of the chemical potential μ, i.e., for varying humidity. We compare results from the capillarity approximation (CA), where Γl is calculated via Eq. (27), with those from DFT, where Γl is calculated via Eq. (29). The results here are for βɛll = 1.2, βɛnn = 0.9, and βɛnl = 1.5. Note that βμ = −3.8 corresponds to a relative humidity of Hr = 80%; βμ = −4 corresponds to Hr = 65%; βμ = −5 corresponds to Hr = 23%; and βμ = −6 corresponds to Hr = 8%. Note also that the CA lines for βμ = −5 and βμ = −6 are barely visible, since the corresponding DFT result is almost identical.

Close modal

In Fig. 5, we show results for the case where βɛnl = 1 (all other parameters are the same as in Fig. 4). This corresponds to a much lower value for the strength of attraction between the nanoparticles and the liquid and, therefore, corresponds to nanoparticles that have a much lower solubility. We see that, as a result of the weaker attraction between the nanoparticles and the liquid, the droplets are, therefore, smaller, as one should expect. Note again the good agreement between the CA and the DFT. Therefore, we conclude that the simple CA is indeed able to capture the effects of varying attraction strengths on determining the equilibrium droplet size.

FIG. 5.

The results here are the same as those in Fig. 4, except now the strength of the attraction between the nanoparticles and the liquid is much lower, with βɛnl = 1; i.e., these are results for nanoparticles with a much lower solubility. Note the change in the range of the vertical axis compared to Fig. 4. In addition, note again that the CA lines for βμ = −5 and βμ = −6 are barely visible since the corresponding DFT result is almost identical.

FIG. 5.

The results here are the same as those in Fig. 4, except now the strength of the attraction between the nanoparticles and the liquid is much lower, with βɛnl = 1; i.e., these are results for nanoparticles with a much lower solubility. Note the change in the range of the vertical axis compared to Fig. 4. In addition, note again that the CA lines for βμ = −5 and βμ = −6 are barely visible since the corresponding DFT result is almost identical.

Close modal

We assume that the non-equilibrium dynamics is governed by dynamical density functional theory (DDFT), as described in detail in Ref. 21. We could include hydrodynamics, e.g., following Ref. 22, but here, we restrict ourselves to the usual overdamped DDFT. DDFT has been derived for Brownian particles suspended in a liquid,13,31 such as the nanoparticles studied here, and has also been developed for molecular liquids.14,32 The latter case is particularly accurate when the fluid is close to equilibrium, which is the case here. As such, DDFT is applicable to both components of the system studied here.

The (lattice) DDFT for the two component system consists of a coupled pair of equations at each lattice site,
(30)
(31)
Note that, in contrast to the DDFT presented in Ref. 21, we have constant mobility coefficients, Ml and Mn, for both the liquid and nanoparticles. It is simple to extend the DDFT to the non-constant case, but we do not find it necessary here. Note also that we set Ml = Mn = 1 for simplicity.

The ensemble-averaged densities at site i, ρil and ρin are now functions of time. The differential operators in (30) and (31) are finite difference approximations on the lattice. Care is needed when applying these to prevent numerical instabilities. We take the approach detailed in Ref. 21, in particular, alternating the direction of the spatial finite difference. Here, we apply this method in both spatial directions, in contrast to Ref. 21, where this was only required in the direction parallel to their wall. For the time evolution, we use an Euler scheme but note that the results are almost indistinguishable from those using higher order schemes such as the fourth order adaptive Runge–Kutta scheme implemented in MATLAB’s ode45.33 

In Figs. 6 and 7, we show the results for two DDFT simulations. Both start from the DFT equilibrium droplet for βɛll = 1.2, βɛnn = 0.9, and βɛnl = 1.5 with βμ = −4.5 (also displayed in Fig. 2). To induce dynamics, we set the liquid density on the boundary of the box to a constant value, corresponding to that of a DFT simulation with a different value of μ. This results in the DDFT simulation equilibrating with the corresponding DFT result. We show two cases, βμ = −3.8 and βμ = −5. The former case, displayed in the two left hand columns of Fig. 6, corresponds to a droplet having moved into a more humid environment than that in which it was initially formed. This results in an increase in the amount of liquid in the box as additional liquid diffuses into the box from the boundary and then condenses onto the droplet, making it grow in size. In contrast, the DDFT results in the two right hand columns of Fig. 6 correspond to the droplet moving to an environment that has a lower humidity than that where it was formed. This is the typical case of aerosol droplets that form in a person’s mouth or respiratory system at a temperature of about 37 °C at 100% relative humidity and then subsequently move out of the mouth into the air, which is a less humid environment. The DDFT results show the size of the droplet decreasing over time as it reduces to a smaller equilibrium size. Note, however, that the droplet never completely evaporates; the nanoparticles within mean that it remains stable, albeit at a smaller size. This can also be seen in Fig. 7, where we plot the total mass of liquid in the simulation box as a function of time for these two cases. The relative errors (in the 1 norm) between the final dynamic profiles and the corresponding equilibria are below 1% in all cases after time 4 × 104; this can be further reduced by running the DDFT simulations for longer times.

FIG. 6.

Snapshots of DDFT simulations, starting from the DFT equilibrium with βμ = −4.5 at times t = 0, t = 4000, and t = 20 000 (top to bottom). Pairs of plots on the left (right) show the DDFT dynamics for the liquid and the nanoparticles when the liquid density on the boundary of the box is set to the value given by the equilibrium DFT computations with βμ = −3.8 (βμ = −5). As expected from the equilibrium calculations, the droplet on the left (right) grows (shrinks) over time.

FIG. 6.

Snapshots of DDFT simulations, starting from the DFT equilibrium with βμ = −4.5 at times t = 0, t = 4000, and t = 20 000 (top to bottom). Pairs of plots on the left (right) show the DDFT dynamics for the liquid and the nanoparticles when the liquid density on the boundary of the box is set to the value given by the equilibrium DFT computations with βμ = −3.8 (βμ = −5). As expected from the equilibrium calculations, the droplet on the left (right) grows (shrinks) over time.

Close modal
FIG. 7.

The mass evolution of the liquid and nanoparticles over time for the DDFT simulations shown in Fig. 6. Dashed lines denote the liquid masses in the corresponding DFT computations. Note, in particular, that increasing (decreasing) βμ from the value of −4.5, which provides the initial condition, causes the amount of liquid in the box to decrease (increase).

FIG. 7.

The mass evolution of the liquid and nanoparticles over time for the DDFT simulations shown in Fig. 6. Dashed lines denote the liquid masses in the corresponding DFT computations. Note, in particular, that increasing (decreasing) βμ from the value of −4.5, which provides the initial condition, causes the amount of liquid in the box to decrease (increase).

Close modal
We note that a crucial aspect of these DDFT simulations relates to the choice of boundary conditions. In the examples presented in Figs. 6 and 7, the nanoparticles are largely concentrated in the middle of the simulation domain, and the interparticle attraction prevents any significant diffusion away from this area. As such, periodic boundary conditions for the nanoparticles are an appropriate choice (any other sensible boundary condition will give almost indistinguishable results). However, the boundary conditions used for the liquid are much more important. In particular, the distance from the surface of the droplet to the edge of the boundary where we keep the liquid density set to the value corresponding to the desired chemical potential value (i.e., relative humidity value specified for each simulation) is all-important. If the box size is increased, i.e., the distance from the droplet to the boundary is increased, then the DDFT simulations take correspondingly longer to equilibrate. What determines the time for a droplet to equilibrate is a combination of two processes: the first relates to the time it takes for liquid to move out of the droplet across the liquid–vapor interface. The second part of the process is that of the liquid diffusing through the vapor surrounding the droplet to reach the boundary and be absorbed. The second process is well understood: for free diffusion from the center to the edge of a circular domain, the total amount in the system Nl(t), given by Eq. (6), follows the well-known result,
(32)
i.e., the amount of liquid decreases exponentially over time with the rate constant λ=D(j1,0/a)2, where D = MlkBT is the diffusion coefficient, j1,0 is the first zero of the Bessel J0(x) function, and a is the radius of the domain. This is the situation where our model reduces to the limit where the densities of the liquid and nanoparticles are small everywhere [where the DDFT equations (30) and (31) reduce to diffusion equations]. However, for the cases of interest here, the additional process of particles crossing the liquid–vapor interface makes the whole equilibration process much slower. In our DDFT simulations, we still observe Nl(t) varying over time with the simple exponential decay form in Eq. (32) (see Fig. 7), but the rate constant λ that we observe is much smaller than the result quoted earlier for the case of simple diffusion from the center to the edge of the domain, due to the additional interface crossing process. Nonetheless, these considerations demonstrate why the distance from the droplet to the edge of the simulation box (i.e., the size we assume for the diffusive boundary layer around the droplets) is important for determining the overall time scale of the equilibration process. That said, we find that as long as the edge of the droplet is 5 or more lattice sites away from the boundary of the box, snapshots over time from simulations in a small box and a larger box are almost indistinguishable.

Our DDFT model can be used to predict the dynamics of aerosol droplets in a great variety of different situations. For example, if we included the external potentials Φil and Φin due to a surface, we could model the slow impact of droplets on a surface and the subsequent spreading and drying process. Illustrative results for the later part of this dynamics can be found, e.g., in Refs. 21 and 23. Here, we restrict ourselves to presenting a pair of illustrative results corresponding to the coalescence of two different sized droplets. In Fig. 8, we show results for droplets joining for the cases when βɛll = 1.2, βɛnn = 0.9, βɛnl = 1.5, and βμ = −5. The initial conditions correspond simply to setting all of the lattice sites within two circular regions to the density values at the center of a single equilibrium droplet for this set of parameter values, while the density outside the circles is set to be that of the corresponding vapor in the single-droplet DFT calculation. The radii of the two circles (i.e., the initial radii of the two droplets) are 5 and 10, with centers at (20,20) and (30,35) for the left hand simulations and (21,21) and (30,35) for the right hand simulations in Fig. 8. Therefore, the only difference between the two simulations is that the smaller droplet is moved slightly closer to the bigger droplet for the right hand set of results. We see, however, that this very small change makes a big difference in the dynamics. In the case on the left, the small droplet shrinks and joins the larger droplet via diffusion through the vapor, while in the case on the right, the whole droplet moves and joins the larger one. Why it is that one sees one process at one distance and the other at a slightly different distance was studied in detail in Ref. 34 in the context of a different DDFT. The mechanism followed by the case on the left is termed joining via the Ostwald mode, which was first described in Refs. 35 and 36 to understand the process of Ostwald ripening, while the mechanism followed by the case on the right is termed the translation mode. One can calculate which mode will dominate by linearizing the DDFT equation around the initial state, and then one obtains two distinct eigenfunctions corresponding to each of these modes. The mode one observes actually occurring is the one with the largest corresponding eigenvalue.34 These results illustrate just one possibility in the hugely complex dynamics of aerosol droplets. Note also that our DDFT model assumes an over-damped (diffusive) dynamics. If we extended our theory to include the effects of inertia, then, for example, the evaporation and coalescence of droplets in the turbulent airflow following a person sneezing could be investigated.37 However, we do not pursue that direction here. The examples in Fig. 8 illustrate that the interplay of dynamics with an underlying complex free energy landscape can result in rather complicated dynamics. In the following section, we illustrate this point further, albeit with examples that correspond to a somewhat unlikely (in nature) initial state.

FIG. 8.

Snapshots from DDFT simulations, starting from two circular distributions at times t = 0, t = 20, t = 1200, and t = 2000 (left, top to bottom) and t = 0, t = 10, t = 100, and t = 1000 (right, top to bottom). The simulations differ only in the initial location of the smaller droplet (see text for details). However, this small difference in initial condition makes a very significant difference in the manner in which the two droplets coalesce.

FIG. 8.

Snapshots from DDFT simulations, starting from two circular distributions at times t = 0, t = 20, t = 1200, and t = 2000 (left, top to bottom) and t = 0, t = 10, t = 100, and t = 1000 (right, top to bottom). The simulations differ only in the initial location of the smaller droplet (see text for details). However, this small difference in initial condition makes a very significant difference in the manner in which the two droplets coalesce.

Close modal

In Fig. 9, we present results corresponding to an initial state where the density of the liquid is uniform throughout the system while the nanoparticles are gathered within a square region in the center of the box. We then perform both the DFT Picard minimization and the DDFT simulation (with periodic boundary conditions for both the fluid and nanoparticles). The initial and final average densities of the liquid in the box are the same for both systems. In the DDFT, the average density is a conserved quantity throughout the dynamics. For the Picard iteration, in this case, it is roughly constant, but more generally, this fictitious dynamics does not preserve mass between iterations. Our Picard iteration results are obtained with the mixing parameter α = 0.01.19,38 Interestingly, the results from these two approaches have very different paths to (the same) equilibrium. This demonstrates that in this case, the “quasi-dynamics” generated by the Picard scheme is not a good approximation of the DDFT dynamics. We see some rather striking transient states during the (realistic) DDFT dynamics displayed on the left of Fig. 9, where the initial square block of nanoparticles breaks up into four smaller droplets that then subsequently re-coalesce into the single final droplet state. This complex dynamics is driven by a competition between bulk and interfacial contributions to the free energy, with each dominating at different stages of the dynamics.

FIG. 9.

Snapshots from a DDFT simulation (left) and from Picard iteration (a fictitious dynamics, right), starting from an initial condition where the liquid density is uniform and the nanoparticles are in a square region at the center of the box. The initial average density of the liquid is selected so that the final states are the same as the equilibrium DFT calculation for βμ = −5. The final equilibria from the two dynamics are the same, but the intermediate states during the evolution are very different. The DDFT profiles on the left are for the times t = 0, 10, 30, 120, and 104. The Picard profiles are from iterations 0, 200, 600, 2400, and 4000.

FIG. 9.

Snapshots from a DDFT simulation (left) and from Picard iteration (a fictitious dynamics, right), starting from an initial condition where the liquid density is uniform and the nanoparticles are in a square region at the center of the box. The initial average density of the liquid is selected so that the final states are the same as the equilibrium DFT calculation for βμ = −5. The final equilibria from the two dynamics are the same, but the intermediate states during the evolution are very different. The DDFT profiles on the left are for the times t = 0, 10, 30, 120, and 104. The Picard profiles are from iterations 0, 200, 600, 2400, and 4000.

Close modal

We have presented a simple capillarity-approximation based theory for the size of nanoparticle laden liquid aerosol droplets. Our theory predicts how the size of the droplets varies depending on the vapor temperature, humidity, the number of nanoparticles within the droplet, and also the nature of the interactions between the nanoparticles and the liquid. We have validated our simple theory by comparing it with results from DFT. Our lattice DFT yields the density distribution of the particles within the aerosol droplets in addition to all the relevant thermodynamic quantities, such as the changes in the liquid–vapor interfacial tension due to varying concentrations of nanoparticles within the droplet. We have also developed a DDFT model able to describe complex dynamical phenomena, such as droplet coalescence. Concrete examples of the types of airborne aerosol systems that our model can be applied to include determining the stability (and, therefore, the lifetime) of exhaled droplets that can lead to the spread of COVID and other diseases, aerosol based therapies such as biomolecule inhalation therapy,39 crop spraying, and the myriad of different aerosols playing important roles in the world’s atmosphere.8–10 

Our simple CA model is useful for quick estimation of droplet sizes as a function of particle loading. For more precise calculations, our CA model could easily be improved by replacing the simple lattice-gas free energy (15) used here with a more accurate equation of state. For example, the Mansoori–Carnahan–Starling–Leland equation of state for hard-sphere mixtures38,40 could easily be used instead, or one of the many other accurate bulk fluid equations of state that are available in the literature (see, e.g., Refs. 41–43). The choice of a particular equation of state would be guided by obtaining additional information about the precise form of the molecular interactions in the system.

For those interested in a more accurate description of the density distribution of the liquid and nanoparticles within droplets, in future work, one could replace the lattice DFT used here with a theory based on the lattice DFT of Refs. 44 and 45, which gives an improved approach for dealing with the nearest-neighbor attractions between particles on a lattice. Alternatively, one could improve the model by using an accurate continuum DFT,12 such as a DFT based on fundamental measure theory.38,46 Such an approach would give a much better description of the liquid structure within droplets. This approach may be needed, especially in cases where particles aggregate at the droplet liquid–vapor interface. For example, size-selectivity can occur in the drying of colloidal films containing two sizes of nanoparticles,47–49 so such effects may occur in the drying of aerosol droplets containing particle mixtures, potentially resulting in highly structured final states.50 

This research was funded by the London Mathematical Society, the International Centre for Mathematical Sciences, and the Loughborough University Institute of Advanced Studies. We are grateful to Emiliano Renzi and David Sibley for their valuable discussions.

The authors have no conflicts to disclose.

A. J. Archer: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Project administration (equal); Writing – original draft (equal); Writing – review & editing (equal). B. D. Goddard: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Project administration (equal); Writing – original draft (equal); Writing – review & editing (equal). R. Roth: Conceptualization (equal); Funding acquisition (equal); Project administration (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
W. F.
Wells
, “
On air-borne infection. Study II. Droplets and droplet nuclei
,”
Am. J. Epidemiol.
20
,
611
618
(
1934
).
2.
X.
Xie
,
Y.
Li
,
A. T. Y.
Chwang
,
P. L.
Ho
, and
W. H.
Seto
, “
How far droplets can move in indoor environments—Revisiting the wells evaporation–falling curve
,”
Indoor Air
17
,
211
(
2007
).
3.
R. R.
Netz
, “
Mechanisms of airborne infection via evaporating and sedimenting droplets produced by speaking
,”
J. Phys. Chem. B
124
,
7093
(
2020
).
4.
C.
Seyfert
,
J.
Rodríguez-Rodríguez
,
D.
Lohse
, and
A.
Marin
, “
Stability of respiratory-like droplets under evaporation
,”
Phys. Rev. Fluids
7
,
023603
(
2022
).
5.
W. C. K.
Poon
,
A. T.
Brown
,
S. O. L.
Direito
,
D. J.
Hodgson
,
L.
Le Nagard
,
A.
Lips
,
C. E.
MacPhee
,
D.
Marenduzzo
,
J. R.
Royer
,
A. F.
Silva
et al, “
Soft matter science and the COVID-19 pandemic
,”
Soft Matter
16
,
8310
(
2020
).
6.
E. P.
Vejerano
and
L. C.
Marr
, “
Physico-chemical characteristics of evaporating respiratory fluid droplets
,”
J. R. Soc., Interface
15
,
20170939
(
2018
).
7.
K. A.
Kormuth
,
K.
Lin
,
A. J.
Prussin
,
E. P.
Vejerano
,
A. J.
Tiwari
,
S. S.
Cox
,
M. M.
Myerburg
,
S. S.
Lakdawala
, and
L. C.
Marr
, “
Influenza virus infectivity is retained in aerosols and droplets independent of relative humidity
,”
J. Infect. Dis.
218
,
739
(
2018
).
8.
B.
Stevens
and
G.
Feingold
, “
Untangling aerosol effects on clouds and precipitation in a buffered system
,”
Nature
461
,
607
(
2009
).
9.
H.
Wei
,
E. P.
Vejerano
,
W.
Leng
,
Q.
Huang
,
M. R.
Willner
,
L. C.
Marr
, and
P. J.
Vikesland
, “
Aerosol microdroplets exhibit a stable pH gradient
,”
Proc. Natl. Acad. Sci. U. S. A.
115
,
7272
(
2018
).
10.
E.
Von Schneidemesser
,
P. S.
Monks
,
J. D.
Allan
,
L.
Bruhwiler
,
P.
Forster
,
D.
Fowler
,
A.
Lauer
,
W. T.
Morgan
,
P.
Paasonen
,
M.
Righi
,
K.
Sindelarova
, and
M. A.
Sutton
, “
Chemistry and the linkages between air quality and climate change
,”
Chem. Rev.
115
,
3856
(
2015
).
11.
R.
Evans
, “
The nature of the liquid-vapour interface and other topics in the statistical mechanics of non-uniform, classical fluids
,”
Adv. Phys.
28
,
143
(
1979
).
12.
J.-P.
Hansen
and
I. R.
McDonald
,
Theory of Simple Liquids: With Applications to Soft Matter
(
Academic Press
,
2013
).
13.
U. M. B.
Marconi
and
P.
Tarazona
, “
Dynamic density functional theory of fluids
,”
J. Chem. Phys.
110
,
8032
(
1999
).
14.
A. J.
Archer
, “
Dynamical density functional theory for dense atomic liquids
,”
J. Phys.: Condens. Matter
18
,
5617
(
2006
).
15.
M.
te Vrugt
,
H.
Löwen
, and
R.
Wittkowski
, “
Classical dynamical density functional theory: From fundamentals to applications
,”
Adv. Phys.
69
,
121
(
2020
).
16.
E.
Kierlik
,
P. A.
Monson
,
M. L.
Rosinberg
,
L.
Sarkisov
, and
G.
Tarjus
, “
Capillary condensation in disordered porous materials: Hysteresis versus equilibrium behavior
,”
Phys. Rev. Lett.
87
,
055701
(
2001
).
17.
H.-J.
Woo
,
L.
Sarkisov
, and
P. A.
Monson
, “
Mean-field theory of fluid adsorption in a porous glass
,”
Langmuir
17
,
7472
(
2001
).
18.
D.
Schneider
,
R.
Valiullin
, and
P. A.
Monson
, “
Filling dynamics of closed end nanocapillaries
,”
Langmuir
30
,
1290
(
2014
).
19.
A. P.
Hughes
,
U.
Thiele
, and
A. J.
Archer
, “
An introduction to inhomogeneous liquids, density functional theory, and the wetting transition
,”
Am. J. Phys.
82
,
1119
(
2014
).
20.
A. P.
Hughes
,
U.
Thiele
, and
A. J.
Archer
, “
Liquid drops on a surface: Using density functional theory to calculate the binding potential and drop profiles and comparing with results from mesoscopic modelling
,”
J. Chem. Phys.
142
,
074702
(
2015
).
21.
C.
Chalmers
,
R.
Smith
, and
A. J.
Archer
, “
Dynamical density functional theory for the evaporation of droplets of nanoparticle suspension
,”
Langmuir
33
,
14490
(
2017
).
22.
M.
Areshi
,
D.
Tseluiko
, and
A. J.
Archer
, “
Kinetic Monte Carlo and hydrodynamic modeling of droplet dynamics on surfaces, including evaporation and condensation
,”
Phys. Rev. Fluids
4
,
104006
(
2019
).
23.
C. M.
Perez
,
M.
Rey
,
B. D.
Goddard
, and
J. H. J.
Thijssen
, “
Changing the flow profile and resulting drying pattern of dispersion droplets via contact angle modification
,” arXiv:2111.00464 (
2021
).
24.
M. J.
Robbins
,
A. J.
Archer
, and
U.
Thiele
, “
Modelling the evaporation of thin films of colloidal suspensions using dynamical density functional theory
,”
J. Phys.: Condens. Matter
23
,
415102
(
2011
).
25.
C.
Chalmers
,
R.
Smith
, and
A. J.
Archer
, “
Modelling the evaporation of nanoparticle suspensions from heterogeneous surfaces
,”
J. Phys.: Condens. Matter
29
,
295102
(
2017
).
26.
C.
Chipot
and
A.
Pohorille
,
Free Energy Calculations
(
Springer
,
2007
), Vol.
86
.
27.
R.
Roth
,
Y.
Harano
, and
M.
Kinoshita
, “
Morphometric approach to the solvation free energy of complex molecules
,”
Phys. Rev. Lett.
97
,
078101
(
2006
).
28.
M. K.
Coe
,
R.
Evans
, and
N. B.
Wilding
, “
Understanding the physics of hydrophobic solvation
,”
J. Chem. Phys.
158
,
034508
(
2023
).
29.
D.
Woywod
and
M.
Schoen
, “
Phase behavior of confined symmetric binary mixtures
,”
Phys. Rev. E
67
,
026122
(
2003
).
30.
M.
Areshi
, “
Mathematical modelling of liquids on surfaces
,” Ph.D. thesis (
Loughborough University
,
2020
).
31.
U. M. B.
Marconi
and
P.
Tarazona
, “
Dynamic density functional theory of fluids
,”
J. Phys.: Condens. Matter
12
,
A413
(
2000
).
32.
A. J.
Archer
, “
Dynamical density functional theory for molecular and colloidal fluids: A microscopic approach to fluid mechanics
,”
J. Chem. Phys.
130
,
014509
(
2009
).
33.
J. R.
Dormand
and
P. J.
Prince
, “
A family of embedded Runge-Kutta formulae
,”
J. Comput. Appl. Math.
6
,
19
(
1980
).
34.
A.
Pototsky
,
U.
Thiele
, and
A. J.
Archer
, “
Coarsening modes of clusters of aggregating particles
,”
Phys. Rev. E
89
,
032144
(
2014
).
35.
I. M.
Lifshitz
and
V. V.
Slyozov
, “
The kinetics of precipitation from supersaturated solid solutions
,”
J. Phys. Chem. Solids
19
,
35
(
1961
).
36.
C.
Wagner
, “
Theorie der alterung von niederschlägen durch umlösen (Ostwald-reifung)
,”
Z. Elektrochem.
65
,
581
(
1961
).
37.
E.
Renzi
and
A.
Clarke
, “
Life of a droplet: Buoyant vortex dynamics drives the fate of micro-particle expiratory ejecta
,”
Phys. Fluids
32
,
123301
(
2020
).
38.
R.
Roth
, “
Fundamental measure theory for hard-sphere mixtures: A review
,”
J. Phys.: Condens. Matter
22
,
063102
(
2010
).
39.
M.
Roudini
,
J.
Manuel Rosselló
,
O.
Manor
,
C.-D.
Ohl
, and
A.
Winkler
, “
Acoustic resonance effects and cavitation in SAW aerosol generation
,”
Ultrason. Sonochem.
98
,
106530
(
2023
).
40.
G. A.
Mansoori
,
N. F.
Carnahan
,
K. E.
Starling
, and
T. W.
Leland
, Jr.
, “
Equilibrium thermodynamic properties of the mixture of hard spheres
,”
J. Chem. Phys.
54
,
1523
(
1971
).
41.
W. G.
Chapman
,
K. E.
Gubbins
,
G.
Jackson
, and
M.
Radosz
, “
SAFT: Equation-of-state solution model for associating fluids
,”
Fluid Phase Equilib.
52
,
31
(
1989
).
42.
G. M.
Kontogeorgis
,
M. L.
Michelsen
,
G. K.
Folas
,
S.
Derawi
,
N.
Von Solms
, and
E. H.
Stenby
, “
Ten years with the CPA (cubic-plus-association) equation of state. Part 2. Cross-associating and multicomponent systems
,”
Ind. Eng. Chem. Res.
45
,
4869
(
2006
).
43.
J.
Gross
and
J.
Vrabec
, “
An equation-of-state contribution for polar components: Dipolar molecules
,”
AIChE J.
52
,
1194
(
2006
).
44.
M.
Maeritz
and
M.
Oettel
, “
Density functional for the lattice gas from fundamental measure theory
,”
Phys. Rev. E
104
,
024124
(
2021
).
45.
M.
Maeritz
and
M.
Oettel
, “
Droplet condensation in the lattice gas with density functional theory
,”
Phys. Rev. E
104
,
034127
(
2021
).
46.
Y.
Rosenfeld
, “
Free-energy model for the inhomogeneous hard-sphere fluid mixture and density-functional theory of freezing
,”
Phys. Rev. Lett.
63
,
980
(
1989
).
47.
A.
Fortini
,
I.
Martín-Fabiani
,
J. L.
De La Haye
,
P.-Y.
Dugas
,
M.
Lansalot
,
F.
D’Agosto
,
E.
Bourgeat-Lami
,
J. L.
Keddie
, and
R. P.
Sear
, “
Dynamic stratification in drying films of colloidal mixtures
,”
Phys. Rev. Lett.
116
,
118301
(
2016
).
48.
B.
He
,
I.
Martín-Fabiani
,
R.
Roth
,
G. I.
Tóth
, and
A. J.
Archer
, “
Dynamical density functional theory for the drying and stratification of binary colloidal dispersions
,”
Langmuir
37
,
1399
(
2021
).
49.
M.
Kundu
and
M. P.
Howard
, “
Dynamic density functional theory for drying colloidal suspensions: Comparison of hard-sphere free-energy functionals
,”
J. Chem. Phys.
157
,
184904
(
2022
).
50.
W.
Liu
,
J.
Midya
,
M.
Kappl
,
H.-J.
Butt
, and
A.
Nikoubashman
, “
Segregation in drying binary colloidal droplets
,”
ACS Nano
13
,
4972
(
2019
).
Published open access through an agreement with JISC Collections