Hostname: page-component-68c7f8b79f-rgmxm Total loading time: 0 Render date: 2025-12-18T05:52:32.670Z Has data issue: false hasContentIssue false

Revisiting droplet combustion: a nearly universal shrinkage kinetic law driven by flame-induced buoyant convection

Published online by Cambridge University Press:  15 December 2025

Chong-An Fang
Affiliation:
Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan
Chao-Yi Yang
Affiliation:
Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan
Shou-Yin Yang
Affiliation:
Department of Power Mechanical Engineering, National Formosa University, Yunlin 632, Taiwan
Hsien-Hung Wei*
Affiliation:
Department of Chemical Engineering, National Cheng Kung University, Tainan 701, Taiwan
*
Corresponding author: Hsien-Hung Wei, hhwei@mail.ncku.edu.tw

Abstract

A burning droplet in normal gravity inevitably encounters buoyant convection set up by the flame, which can significantly impact its shrinkage kinetics traditionally described by the D2-law. However, the detailed mechanism governing droplet vapourisation under such self-generated flame-driven buoyant convection remains elusive. Here, we present both experimental and theoretical evidence highlighting the critical role of buoyant convection in droplet combustion. Experimentally, we precisely measure the values of the shrinkage exponent n for various liquid fuels, revealing a significant departure from the D2-law. While the measured n values consistently fall within the narrow range 2.6–2.7, they exhibit a slight increase with the fuel’s boiling point. A more general and in-depth theory is also developed to explain such small but systematic variations, revealing that differences in flow and thermal boundary layer structures – arising from varying combustion intensities – may account for the observed trends. Our theory predicts three distinct values of n, namely 2.6, 8/3 ≈ 2.67 and 35/13 ≈ 2.69, successfully capturing slight differences in n among various fuels. This is the first study demonstrating that the shrinkage kinetics in droplet vapourisation driven by flame-induced buoyant convection is nearly universal, determined solely by the underlying transport mechanisms, although these can be significantly altered due to their high susceptibility to detailed fuel chemistry and combustion kinetics. The present theoretical framework not only enables accurate prediction and control of burning droplet behaviour, but also is extendable to analyse more complex combustion processes involving a broader range of fuel types and flow conditions.

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

Fuel combustion is a widely used energy generation process, capable of producing immense power or heat through rapid thermal oxidation, as observed in engines, boilers and furnaces. As a fuel is typically in liquid form, its combustion is often achieved by atomising it into a cloud of fine droplets, which increases the surface area for interaction with gaseous oxidisers, resulting in a much greater heat release (Sirignano Reference Sirignano1983). Since the burning efficiency of a fuel droplet is often closely linked to its vapourisation rate, and relies on how the burning is performed, understanding fuel droplet vapourisation under various conditions is essential for optimising combustion efficiency (Téré et al. Reference Téré, Christian, Kayaba, Boubou, Sayouba, Tizane, Jean, Oumar, Belkacem and Antoine2022; Wang et al. Reference Wang, Yuan, Huang, Cao, Yuzhou Wang and Cheng2022).

The burning performance of a fuel is often evaluated by how quickly it releases heat by observing how fast a fuel droplet reduces its size or how long its lifetime spans. The D 2-law is widely used relation for this purpose, stating that the square of the instantaneous diameter D of a vapourising droplet (having initial diameter D 0) decreases linearly with time t (Godsave Reference Godsave1953; Faeth Reference Faeth1977; Law Reference Law1982):

(1.1) \begin{equation} D^{2}=D_{0}^{2}-Kt, \end{equation}

where K is the burning rate constant usually made of the properties of a liquid fuel and applied ambient conditions, measuring the burning potential of the fuel. While the D 2-law was originally derived for pure droplet evaporation in the absence of gravity without convection (Spalding Reference Spalding1953; Dalla Barba et al. Reference Dalla Barba, Wang and Picano2021), it is somewhat surprising that it can still reasonably approximate the shrinking behaviour of a burning droplet in a normal gravity environment where natural convection prevails (Sato et al. Reference Sato, Tsue, Niwa and Kono1990; Yadev et al. Reference Yadev, Chowdhury and Srivastava2017). Even in more complex scenarios involving multi-component or blended fuels (Chow et al. Reference Chow, Ooi, Chee, Pun, Tran, Leong and Lim2021; Li et al. Reference Li, Tian, Han, Bao, Meng and Lin2021), emulsions (Priyadarshini, Kushari & Rao Reference Priyadarshini, Kushari and Rao2024), and the addition of particles (Ferrão et al. Reference Ferrão, Mendes, Mendes, Moita and Silva2025; Parveg & Ratner Reference Parveg and Ratner2025), the D 2-law appears to effectively describe the observed droplet size evolutions. It is remarkable that the D 2-law can approximate droplet shrinkage behaviour across a wide range of systems and conditions, despite extending beyond its original theoretical applicability.

A closer inspection reveals that such seemingly universal behaviour described by the D 2-law may be misleading. To illustrate this, consider a generalised shrinkage kinetics model following a power-law form:

(1.2) \begin{equation} D^{n}=D_{0}^{n}-Kt. \end{equation}

By plotting D 2 against t with different values of the shrinkage exponent n, it becomes evident (figure 1 a) that the profiles still look approximately linear, resembling the D 2-law (1.1), even when n deviates slightly from 2. When n< 2, the curve appears slightly concave, whereas for n> 2, it becomes somewhat convex. The latter behaviour is frequently observed in droplet combustion experiments under gravity, as illustrated in figure 1(b) using experimental data. This strongly suggests that the D 2-law may not be the true governing law for a burning droplet in a gravitational environment.

Figure 1. Illustration of the D 2t plot with different values of the shrinkage exponent n in (a), showing that the profiles with n slightly different from 2 can still appear, resembling the classical D 2-law. For n< 2 (e.g. n = 1.5), the curve becomes slightly concave, while for n> 2 (e.g. n = 2.5), it exhibits a mildly convex shape. Such convex profiles can occur in droplet combustion under gravity, as shown in (b) with experimental data for tetradecane, ethanol and kerosene in the present work.

The present work is to examine the impact of inevitable buoyant convection set up by burning under gravity in droplet combustion processes. Such buoyant convection can give rise to simultaneous momentum, heat and mass transfer, markedly different from the convection-free situation described by the D 2-law. Previous studies have attempted to account for natural convection either by applying the D 2-law to determine the vapourisation rate constant (Sato et al Reference Sato, Tsue, Niwa and Kono1990; Verwey & Birouk Reference Verwey and Birouk2018; Murakami et al. Reference Murakami, Nomura and Suganuma2021; Ferrão et al. Reference Ferrão, Mendes, Mendes, Moita and Silva2025) or by employing Ranz–Marshall-type correlations for the heat transfer coefficient across the droplet surface (Ranz & Marshall Reference Ranz and Marshall1952; Sazhin Reference Sazhin2017; Wang et al. Reference Wang, Cui, Bi, Liu, Dong, Xing, Li and Liu2020a ; Téré et al. Reference Téré, Christian, Kayaba, Boubou, Sayouba, Tizane, Jean, Oumar, Belkacem and Antoine2022). The Reynolds numbers or the Grashof numbers in these attempts are still based on droplet size. While this is appropriate for droplet evaporation, it may be questionable for droplet combustion because it overlooks the fact that convection is driven by the flame rather than the droplet itself – the flame width should be the more relevant length scale. Numerical simulations have been conducted to fully couple momentum, heat and mass transfer induced by buoyant convection (Saufi et al. Reference Saufi, Frassoldati, Faravelli and Cuoci2021). However, since the resulting droplet size evolution is presented in $D^2{-}t$ plots, whether the effects modify the shrinkage kinetic law or whether it significantly departs from the D 2-law is not clear. It has been shown that a non-square-diameter law can be derived for an evaporating droplet subjected to forced convection and is able to successfully capture the shrinking behaviour of the droplet (Starinskaya et al. Reference Starinskaya, Miskiv, Nazarov, Terekhov, Terekhov, Rybdylova and Sazhin2021). A similar result has also been observed for an evaporating sessile droplet in the presence of natural convection (Misyura Reference Misyura2018). Based on this, it is conceivable that a distinct non-square-diameter law should also govern the shrinking behaviour of a burning droplet influenced by buoyant convection generated by the flame.

Another reason why the classical D 2-law cannot represent the true burning kinetic law for droplets under gravity stems purely from dimensional considerations. The presence of gravity fundamentally alters how the relevant length scale evolves with time. A familiar example is the contrast between free fall under gravitational acceleration g and motion at constant velocity V: the distance travelled under gravity is gt 2/2, while for constant velocity it is Vt. By analogy, we expect a similar shift in droplet combustion behaviour. The D 2-law, which applies in the absence of gravity, must transition to a Dn -law when gravity is present, with an exponent n that differs from 2.

The sign of a possible violation of the D 2-law for droplet combustion under gravity has been implied in the study by Law & Williams (Reference Law and Williams1972). They did notice that buoyant convection can affect the burning rate constant K in the D 2-law (1.1), and proposed an empirical correction for K by taking into account such convection in terms of the Grashof number. In their study, they used –d(D 2)/dt to measure the value of K that constantly changes with time t. This approach implies that the droplet shrinkage kinetics has departed from the D 2-law, but assuming that the impacts of the power-law change can be absorbed into K. Similar ideas and approaches are also employed by Verwey & Birouk (Reference Verwey and Birouk2018), who examined effects of natural convection on droplet evaporation. These early studies imply that there may exist some value different from 2 for the exponent n in K = –d(D n )/dt or D n –t plot under which K is constant. The value of K determined in this way no longer possesses units mm2s–1 like those in the D 2-law, but turns into mm n s–1, which reflects the true impacts arising from buoyant convection. So we believe that a different non-square-diameter law should exist for a burning droplet due to buoyant convection.

Motivated by the above, it is essential to identify the shrinkage kinetic law in droplet combustion under the influence of buoyant convection in a normal gravity environment. Chen et al. (Reference Chen, Yang, Yang and Wei2024) made the first attempt in this direction by conducting droplet combustion experiments using the suspended fibre technique (Chauveau et al. Reference Chauveau, Birouk, Halter and Gökalp2019; Wang et al. Reference Wang, Cui, Bi, Liu, Dong, Xing, Li and Liu2020). In their study, the value of the shrinkage exponent n for a given fuel droplet is determined from the slope in the plot of (1– t/t life ) against D/D 0 in log–log scale, where t life is the lifetime of the droplet. Using this dynamic slope approach, they found that while the D 2-law can be recovered for droplet evaporation (see figure 2 a), the values of n for droplet combustion across a variety of pure liquid fuels are approximately 2.6, indicating a significant departure from the D 2-law. In fact, this dynamic slope approach has been proven quite robust, as it can be applied to determine the values of n for a wide range of droplet vapourisation processes (Fang et al. Reference Fang, Tseng, Yang and Wei2026).

Figure 2. Plots of (1 – t/t life ) versus D/D 0 for measuring the shrinkage exponent n during ethanol droplet evaporation (based on 10 realisations), using (a) the suspended method (fibre diameter ∼ 35 μm) and (b) the cross-fibre method (fibre diameter ∼ 100 μm). The measured n value in (a) is pretty close to the ideal value of 2, indicating that the supporting fibre has minimal influence and the D 2-law holds. In contrast, (b) shows a significantly larger n, suggesting that the presence of cross-fibres notably affects the evaporation process.

The dynamic slope approach adopted by Chen et al. (Reference Chen, Yang, Yang and Wei2024) still has an inherent issue that the droplet lifetime t life inevitably contains contributions from the support fibres. This may cause n in (1.2) to deviate from the true value in the absence of fibres. While this does not appear to be an issue in droplet evaporation in the presence of a suspended fibre under which the D 2-law can still hold (Chen et al. Reference Chen, Yang, Yang and Wei2024), it is not clear whether the observed departure from the D 2-law in droplet combustion is genuine or merely an artefact of the fibre. This concern arises because as the droplet the approaches the end of vapourisation and nearly dries out, its size becomes comparable to that of the fibre. This may escalate the fibre heating, potentially overshadowing the ambient heating from the surrounding gas, particularly under elevated temperature conditions (Wang et al. Reference Wang, Huang, Qiao, Ju and Sun2020b ). In fact, the fibre here acts as an additional heat source, which may cause n to increase towards 3. It is actually observed in droplet evaporation experiments, using both the suspended fibre method and the cross-fibre method, that the measured value of n for the former is close to the ideal value 2, whereas that for the latter significantly exceeds 2, as illustrated in figure 2(b). This suggests that the effects of the fibre may obscure the true nature of a droplet vapourisation process. For droplet combustion, this situation is undesirable since it is crucial for the measured burning rate constant K of a liquid fuel to accurately reflect the fuel’s ability to burn, free from the influence of support fibres. Therefore, to extract the true value of K from experimental data, it is essential not only to determine the correct shrinkage kinetic law for a burning droplet, but also to ensure that this law is not significantly affected by the presence of support fibres.

The aforementioned issue necessitates a more reliable method to measure the value of n for droplet combustion under gravity while minimising the influence of support fibres. In this work, we will employ a thinner fibre with a diameter smaller than that used by Chen et al. (Reference Chen, Yang, Yang and Wei2024) and use the same suspended fibre technique for our experiments. To further mitigate near-contact interferences from the fibre, we will determine the value of n by directly fitting experimental data to (1.2), using data points that are sufficiently away from the end of vapourisation. If the measured values of n are found to be consistent with those obtained from the dynamic slope approach, then we can confidently conclude that the observed shrinkage power law is authentic and not affected by fibre interference. As will be shown in § 3, this is indeed the case.

Beyond experimental investigation, we also seek a fundamental understanding of why droplet combustion exhibits a positive departure from the D 2-law due to buoyant convection driven by burning. This departure is not immediately obvious, as convection typically reduces the shrinkage exponent n below 2 (Yuge Reference Yuge1960; Starinskaya et al. Reference Starinskaya, Miskiv, Nazarov, Terekhov, Terekhov, Rybdylova and Sazhin2021).

Recently, we proposed a theory showing that the observed departure can be well captured by a new D 8/3-law, derived purely from transport effects induced by flame-driven buoyant convection, independent of detailed combustion reaction kinetics (Chen et al. Reference Chen, Yang, Yang and Wei2024). However, a closer inspection – particularly when minimising the impact of the support fibre – reveals that the measured values of n actually systematically deviate from the theoretical value 8/3, with variations correlated to fuel volatility. Given that flame behaviour also differs across fuels, we suspect that these variations influence the underlying transport mechanisms, prompting us to explore ways to reconcile the small discrepancy between experiment and theory. As will be demonstrated in § 4, slightly differences in n among various fuels can be attributed to different degrees of burning, which, in turn, lead to distinct flow and thermal boundary layer structures around the droplet. These structural variations, which arise from different fuel chemistry, affect the dependence of the heat transfer coefficient on droplet radius, ultimately accounting for the observed deviations in n.

In light of the above, our goals are not only to confirm the true shrinkage kinetics of a burning droplet, free from the influence of the support fibre, but also to investigate the detailed mechanisms behind the small but systematic variations in the measured n values in connection to fuel chemistry and combustion kinetics. Prior to presenting our results, we first explain how we set up the experiment and handle the data.

2. Experiment and data analysis methods

2.1. Experimental set-up and procedures

Figure 3 is the schematic diagram for our experimental set-up. The chamber is made of four vertical aluminium framing posts, with glass panels covering the front and rear sides. The top and bottom sides of the chamber remain open to ensure proper airflow. The surrounding environment is maintained at 25 °C under air-conditioned conditions, with relative humidity 50 %.

Figure 3. Schematic diagram of the experimental set-up used in the present droplet combustion study.

To support the droplet, we use a low thermal conductive ceramic fibre to minimise heat transfer from the fibre to the droplet (Lai & Pan Reference Lai and Pan2023). The fibre has thermal conductivity 0.12 W m–1. Two fibre diameters are employed: 35 μm and 100 μm. The fibre is suspended from the top of the chamber, with one end attached on a sticky tape across the diagonal extrusions. To prevent the droplet sliding down along the fibre, we tie a small knot (∼140 μm in size) at the end of the fibre. Before each experiment, a droplet of the desired liquid is carefully placed onto the knot using a syringe needle. The starting diameter of the droplet is 0.5–1.0 mm.

After igniting the droplet, we monitor and record its burning and shrinkage using a high-speed CCD camera (Phantom v7.3) positioned in front of the chamber. The video shooting is undertaken at 1000 frames per second, with exposure time 1 ms. We further add an LED light source (LEDD1B, 14.4W) behind the rear side of the chamber. This is to ensure that the recording is made in a backlighting manner. Once recorded, the video is stored and converted into a sequence of images with resolution 800 × 600 pixels. The total number of the images is 500–3000, depending on the droplet lifetime. These images are then post-processed using MATLAB for the subsequent data acquisition and analysis.

In the experiment, we select a range of common pure liquid fuels as well as practically used fuel blends. For each fuel, we run its burning 50 times. The initial diameter D 0 of a droplet is chosen at a redefined starting moment t = 0 after the early-time heat-up period during which the droplet undergoes a simple heating prior to combustion. To further avoid possible near-contact interferences from the fibre, we use only the data from D = D 0 down to D = 10 % D 0 larger than the fibre diameter. We additionally carry out the corresponding droplet evaporation experiments (10 times) at elevated temperature 70 °C as a control case, confirming that the D 2-law still holds and is not influenced by the presence of the fibre.

Since our goal is to determine the shrinkage exponent n from the measured instantaneous diameter D of a droplet, it is crucial to ensure accurate measurement of D. The methodology for this measurement is described next.

2.2. Using a volume-equivalent sphere to determine the effective droplet diameter

The presence of the fibre causes a droplet to deviate from a perfect spherical shape (see figure 2 a). Accurately determining the instantaneous droplet diameter D from images is critical, as it directly impacts the measured value of n in (1.2). Taking the D 2-law as an example, one may have an impression from (1.1) that D 2 represents the area of the droplet, and hence take D to be the diameter of the projection area of the droplet (Shang et al. Reference Shang, Yang, Xuan, He and Cao2020; Wang et al. Reference Wang, Huang, Qiao, Ju and Sun2020b ). However, this approach can introduce a significant error in determining n, as it neglects the fact that the droplet shrinkage occurs in three dimensions rather than two. Furthermore, if n deviates from 2, then determining D would depend on the specific shrinkage kinetics. This creates a paradox, as n is initially unknown and is precisely what we aim to determine.

A rational approach to measuring D lies in the fact that the shrinkage of a burning droplet results from the consumption of its volume at a rate $\dot{V}$ that is accompanied by the energy required for vapourisation, ρ L $\dot{V}$ $\Delta H$ $ _{v\textit{ap}} $ , where $\Delta H$ $ _{v\textit{ap}} $ is the latent heat of vapourisation, and ρ L is the density of the liquid fuel. Since it is the volume change that drives the phase transition, it is essential to use the droplet’s volume to determine its effective diameter, especially when the droplet deforms due to the fibre.

In order to more accurately measure the shrinkage exponent n, we adopt the volume-equivalent-sphere approach (Nomura et al. Reference Nomura, Ujiie, Rath, Sato and Kono1996; Han et al. Reference Han, Yang, Zhao, Fu, Ma and Song2016), in which the effective diameter D of a deformed droplet is determined by treating the droplet as a sphere having the same volume as the actual droplet. To achieve this, we assume that the droplet profile is axisymmetric, with the interface position r(z) measured from the axis z of symmetry. We perform a volume integral to obtain the volume for this axisymmetric body, representing the droplet’s volume after subtracting the volume of the knot (with diameter d knot ) at the end of the fibre. By equating the net volume to that of the equivalent sphere, we can determine the effective diameter D of the droplet according to

(2.1) \begin{equation} V=\pi \int r^{2}\,{\rm d}z-\frac{\pi }{6}d_{knot}^{3}=\frac{\pi }{6}D^{3}. \end{equation}

It is important to note that subtracting the knot’s volume is necessary because it guarantees D to be identically zero when the droplet has completely vapourised at the end of its burning.

It has been shown that this volume-equivalent-sphere approach is able to effectively address the droplet deformation issue, and successfully recovers the D 2-law (Mandal & Bakshi Reference Mandal and Bakshi2012; Chen et al. Reference Chen, Yang, Yang and Wei2024). In the study of Chen et al. (Reference Chen, Yang, Yang and Wei2024), the exponent n is determined by the dynamic slope method. Since this method involves the droplet lifetime t life , which can be influenced by the presence of the fibre, it remains unclear whether this influence significantly affects the measured values of n in droplet combustion – where fibre heating effects may become more pronounced at elevated temperatures – even though such effects do not appear to impact the D 2-law for droplet evaporation. To address this concern, in addition to using a thinner fibre, we also determine the value of n by directly fitting the data to (1.2), as described next.

2.3. Determining the shrinkage exponent through direct data fitting

To fit the measured D–t data to (1.2), we start with the alternative form

(2.2) \begin{equation} \left(D/D_{0}\right)^{n}=1-K't, \end{equation}

where K = K/D 0 n is the modified burning rate constant, which is also unknown. Therefore, we have two unknown parameters, n and K´, that need to be determined.

The purpose of directly fitting data to (2.2) is to minimise the influence of the fibre as much as possible. This is achieved by selecting the data sufficiently far from the end of the burning. In other words, instead of fitting the entire data set all the way to t life , we select the data points only up to D = D f at a specific moment t f not close to t life . At this moment, (2.2) must satisfy

(2.3) \begin{equation} \left(D_{\kern-1pt f}/D_{0}\right)^{n}=1-K't_{\kern-1pt f}. \end{equation}

The ratio between (2.2) and (2.3) yields

(2.4) \begin{equation} D/D_{\kern-1pt f}=\left[1+G\left(t_{\kern-1pt f}-t\right)\right]^{s}, \end{equation}

where the parameters are now transformed to

(2.5a,b) \begin{equation} s=1/n,\quad G=K'/(1-K't_{\kern-1pt f}). \end{equation}

Now we define Y = ln(D/D f ) and T = t f t. Instead of directly fitting the data to (1.2), we perform a regression of the measured experimental data Y exp = ln (D/D f ) to the model

(2.6) \begin{equation} Y=s\ln (1+G T), \end{equation}

with the new parameters s and G replacing n and K .

The advantage of the recast model above is that the selected data range does not have to start from the initial point D = D 0 at t = 0 – it can start at an arbitrary time, and end at any later moment. This flexibility allows for the exclusion of both the early heat-up period and the late-time period during which the fibre effects become more pronounced. Additionally, this approach offers the benefit of performing a consistency check to verify whether different data ranges still conform to the same kinetic law.

The regression is performed using the least square method for minimising the sum of squared errors, E, between the experimental value Y i exp = ln [D(T i )/D f ] and the model value Y i = Y(T i ) at T i (i = 1 … N) across all points in the selected data set:

(2.7) \begin{equation} E=\sum _{i=1}^{N}\left(Y_{i}^{\textit{exp} }-Y_{i}\right)^{2}. \end{equation}

The error minimisation is carried out under ∂E/∂s = 0 and ∂E/∂G = 0, giving

(2.8a,b) \begin{equation} s=\frac{\sum _{i=1}^{N}Y_{i}^{\textit{exp} }\ln \left(1+GT_{i}\right)}{\sum _{i=1}^{N}\left[\ln \left(1+GT_{i}\right)\right]^{2}},\quad s=\frac{\sum _{i=1}^{N}{Y_{i}^{\textit{exp} }}/({1+GT_{i}})}{\sum _{i=1}^{N}{\ln (1+GT_{i})}/({1+GT_{i}})}. \end{equation}

Combining the above equations gives an equation for G. Once G is found by numerically solving the combined equation, the value of n can be easily determined from s = 1/n by substituting the value of G into (2.8a ) or (2.8b ).

Using the volume-equivalent-sphere approach described in § 2.2 and the direct fitting method outlined above, we can later demonstrate that evaporation of a fibre-suspended droplet still follows the classical D 2-law very well, thereby validating the reliability of our measurement technique in capturing the overall shrinkage kinetics. In contrast, for droplet combustion, the measured n values across different fuels will be shown to consistently fall in a narrow range significantly greater than 2 with little variation, largely independent of fibre diameter or fibre support method. This consistency strongly indicates that the observed deviation from the D 2-law must arise from droplet combustion, rather than from the fibre’s presence or associated deformation effects.

3. Experimental observations

We first present sequential images for ethanol, tetradecane and kerosene during their droplet combustion experiments, as shown in figure 4. We select these fuels because they represent three distinct flame appearances observed in experiments. For highly volatile ethanol (boiling temperature T b = 78 °C), its flame is blue and takes on a raindrop shape (figure 4 a), indicating vigorous burning. In contrast, for less volatile teradecane (T b = 253 °C), the flame remains blue at the bottom, but transitions to red and becomes more elongated in the upper part due to less intense combustion (figure 4 b). For kerosene (T b ∼ 172 °C), where incomplete burning is more pronounced, the flame appears predominately red and extends even further (figure 4 c). Disruptive phenomena such as puffing or micro-explosions are not observed, as our experiments use single-component fuels.

Figure 4. Sequential images showing the droplet combustion processes of (a) ethanol, (b) tetradecane and (c) kerosene. The lower images in each set provide zoomed-in views of the droplets located in the lower portions of the corresponding frames.

Figure 5(a) presents the corresponding log-log plots of (1 – t/t life ) against D/D 0 for these fuels. The slopes yield the values of the shrinkage exponent n = 2.56 ± 0.16, 2.66 ± 0.21 and 2.73 ± 0.22 for ethanol, tetradecane and kerosene, respectively. These values appear to correlate with different flame appearances observed in figure 4, which will be discussed in more detail in § 4.2. In comparison, droplet evaporation experiments are also carried out for these fuels, showing that their shrinkage kinetics closely follows the D 2-law, as shown in figure 5(b).

Figure 5. Plots of (1 − t/t life ) against D/D 0 for determining the values of the shrinkage exponent n for (a) droplet combustion and (b) evaporation experiments using ethanol, tetradecane and kerosene shown in figure 4.

We also determine the values of n for the aforementioned and additional fuels by fitting the data directly to (1.2). The results are presented in figure 6, which plots the measured n values against the boiling point T b . First, we observe that the measured n values remain relatively constant across different fuels, showing little sensitivity to hydrocarbon composition or chemical structure. This suggests that the droplet shrinkage kinetics is governed primarily by universal transport mechanisms, rather than by fuel chemistry or detailed combustion kinetics. In addition, there is essentially no difference between the measured n values through directly fitting the data to (2.2) and those using the dynamic slope method, as shown in figure 6(a). A similar trend is also observed for droplets burning on thicker fibres (100 μm in diameter), although the measured n values are slightly lower, as shown in figure 6(b). All measured values fall within a narrow range approximately 2.6–2.7, clearly deviating from the D 2-law. Together with the re-affirmation of the D 2-law in the corresponding droplet evaporation experiments with fibres (see figure 6 a), the consistency of these findings – across different data analysis approaches and fibre support configurations – strongly indicates that the observed deviation from the D 2-law is not an artefact introduced by the presence of the fibre.

Figure 6. Measured values of the shrinkage exponent n plotted against the boiling points of various liquid fuels. Both droplet combustion (50 realisations) and evaporation (10 realisations) experiments were conducted under gravity for each fuel. (a) In droplet combustion using the suspended fibre technique (fibre diameter ∼ 35 μm), the values of n obtained by the best-fit method (solid triangles) and the dynamic slope method (open triangles) are approximately 2.6–2.7, consistent with the theoretical prediction $8/3$ . These values are notably higher than those approximately 2 observed in the corresponding evaporation experiments (based on the dynamic slope method). (b) Droplet combustion using the cross-fibre technique (fibre diameter ∼ 100 μm) yields similar n values (based on the dynamic slope method), strongly suggesting that the observed departure from the D 2-law arises from combustion effects rather than the presence of the fibres.

It is worth noting that the present droplet evaporation experiments are conducted at an elevated temperature, where natural convection may arise. However, the measured n values for a variety of fuels remain consistently close to 2, indicating that the classical D2 -law still holds under such conditions. This suggests that natural convection effects during evaporation are either negligible or do not significantly alter the shrinkage kinetics, thereby reinforcing the validity of the D2 -law under the tested conditions. In fact, the Grashof number Gr = ΔTR 3/ν 2 here is approximately 6 for a droplet with radius R ∼ 1 mm, subject to a temperature difference ΔT ≈ 45 °C in air, with kinematic viscosity ν = 0.15 cm2 s–1 and thermal expansion coefficient β = 1/T ≈ 0.003 K–1 at ambient temperature T ≈ 343 K. This value of Gr is considered small, indicating that viscous force dominates over buoyancy force, resulting in weak natural convection flow and thereby justifying its negligible contribution to this evaporation process.

Since figure 6 presents combined results obtained using different fibre diameters and various methods for determining n, figure 7 replots the data using only the direct fitting method, which is considered more reliable. As shown in figure 7(a) for thinner fibres (35 μm in diameter), while the measured n values are generally close to the theoretical value n = 8/3 predicted by Chen et al. (Reference Chen, Yang, Yang and Wei2024), a closer inspection reveals a slight increase in n with increasing boiling point T b , which is positively correlated with hydrocarbon content. This trend also persists for thicker fibres (100 μm in diameter), although the measured n values are slightly lower, as shown in figure 7(b). The fact that this trend appears regardless of fibre diameter suggests that it is unlikely to result from fibre effects but rather originates from the intrinsic combustion characteristics of the fuel. This also suggests that the rapid decline near the end of combustion, as shown in figure 1(b), cannot be caused by potential fibre heating effects. This is because the final stage has been excluded from our data fitting analysis used to determine the exponent n, and the measured n values for droplet evaporation remain consistently close to 2 across various fuels – confirming the validity of the D 2-law, and indicating no influence from the fibre. Therefore, the significantly larger n values observed during droplet combustion cannot be attributed to the presence of the fibre. A more detailed justification for the negligible impact of fibre heating is provided in § 4.2.

Figure 7. Effects of the fibre diameter (d f ) on the shrinkage exponent n for various liquid fuels based on the suspended fibre method. The results are obtained via the direct data fitting method, with data extracted sufficiently away from the fibre to minimise interference. This analysis is to reveal systematic variations of n relative to the theoretical value n = 8/3 predicted by Chen et al. (Reference Chen, Yang, Yang and Wei2024), particularly for low-boiling-point fuels (e.g. hexane and ethanol) and practically used fuels (e.g. diesel and kerosene).

It is worth noting that although the measured n values across a variety of fuels fall within the narrow range 2.6–2.7, the differences between fuels are statistically significant owing to the large number of experimental runs (50 per fuel). For example, the representative fuels – highly volatile ethanol (n = 2.56 ± 0.16), moderately volatile tetradecane (n = 2.66 ± 0.21), and low-volatility kerosene (n = 2.73 ± 0.22) – show meaningful distinctions. Comparing ethanol and tetradecane, the t-test yields t = –2.68 and p = 0.0086. A t-test comparing ethanol and kerosene yields t = –4.43 and p = 0.00002. Both p-values are well below 0.01, confirming that the differences in n between fuels are statistically significant.

Along the above lines, figure 8 shows that the measured n values can be approximately categorised into three groups, based on the degree of deviation from the theoretical prediction n = 8/3 by Chen et al. (Reference Chen, Yang, Yang and Wei2024). For volatile fuels such as ethanol and hexane having low values of T b , figure 8(a) shows that the measured n values are approximately 2.6, noticeably lower than the theoretical prediction n = 8/3. For such fuels, their flames are mostly blue (see figure 4 a), indicating that the burning is quite complete and vigorous. For less volatile fuels such as tetradecane (C14H30) and dodecanol (C12H25OH), having higher values of T b , the measured n values increase slightly to 2.65, aligning more closely with the predicted value n = 8/3, as shown in figure 8(b). The flames of these fuels exhibit blue colouration only in the lower regions, while the upper portions turn red due to partially incomplete combustion (see figure 4 b). However, for practically used fuels such as diesel and kerosene, figure 8(c) indicates that the measured n values further increase to 2.7 or higher, systematically exceeding the theoretical prediction n = 8/3. Although their T b values are not as high as those of less volatile pure liquid fuels, their flames appear predominantly red due to severe incomplete combustion (see figure 4 c).

Figure 8. Measured n values may show slight deviations from the theoretical value n = 8/3 predicted by Chen et al. (Reference Chen, Yang, Yang and Wei2024), depending on fuel volatility reflected by boiling point. (a) For more volatile fuels such as ethanol and hexane, their n values are slightly lower than n = 8/3. (b) For less volatile fuels such as tetradecane (C14H30) and dodecanol (C12H25OH), their n values align closely with n = 8/3. (c) For practically used fuels such as diesel and kerosene, their n values become even larger.

4. Comprehensive theory and comparison with experiment

4.1. Droplet vapourisation due to flame-driven buoyant convection

As demonstrated in the preceding section, we have shown that flame-driven droplet combustion exhibits a significant departure from the D 2-law, with the extent of deviation being influenced by fuel volatility. In this section, we present a comprehensive theory to explain these findings.

First and foremost, why n deviates from 2 has to be traced back to the method used to determine n, which is derived from a heat balance over a vapourising fuel droplet of radius R. Assuming that the droplet is rapidly mixed internally, the temperature within it can be considered uniform, as if it were an infinitely conductive sphere (Aggarwal, Tong & Sirignano Reference Aggarwal, Tong and Sirignano1984). Under this assumption, the heat required to vapourise the droplet (with density ρ L and volume V = 4πR 3/3) is primarily supplied from the surrounding gas, driven by the temperature difference ΔT between the flame and the droplet surface (of area A = 4πR 2), as illustrated in figure 9. The corresponding energy balance can be expressed as

(4.1) \begin{equation} \rho _{L}\dot{V}{\Delta} H_{\textit{vap}}=-h\,{\Delta} \textit{TA}. \end{equation}

Figure 9. Schematic figures for different local flow and boundary layer structures around a burning droplet to account for variations of n observed in the experiments. (a) Slipping flow with little soot particle contamination on the droplet surface, giving n = 8/3 ≈ 2.67. (b) Shear flow with severe soot particle contamination on the droplet surface, which results in n = 35/13 ≈ 2.69. (c) Straining flow resulting from strong airflow impingent arising from vigorous flame burning, yielding n = 2.6.

Here, the vapourisation energy provided by the latent heat ΔH vap on the left-hand side is associated with the mass loss of the droplet ρ L $\dot{V}$ = 4πR 2 $\dot{R}$ ρ L (< 0). The heat transfer coefficient h can be evaluated from the Nusselt number, which depends on the detailed thermal boundary layer structure around the droplet:

(4.2) \begin{equation} \textit{Nu}\equiv \textit{hR}/k, \end{equation}

where k is the thermal conductivity of the gas phase. With (4.2), (4.1) can be rewritten as

(4.3) \begin{equation} \rho _{L}{\Delta} H_{\textit{vap}}\frac{{\rm d}R^{2}}{{\rm d}t}=-2k\, \textit{Nu}\,{\Delta} T. \end{equation}

If the heat transfer is purely by conduction, i.e. Nu = 1, then (4.3) reduces to

(4.4a) \begin{equation} \frac{{\rm d}R^{2}}{{\rm d}t}=-2k{\Delta} T/{\Delta} H_{\textit{vap}}\rho _{L}, \end{equation}

which recovers the classical D 2-law given in (1.1). The corresponding vapourisation rate constant is

(4.4b) \begin{equation} K=8\alpha \left(\rho /\rho _{L}\right) \left(C_{p}{\Delta} T/{\Delta} H_{\textit{vap}}\right), \end{equation}

where α = k/ρC p is the thermal diffusivity of the gas phase (with gas density ρ and heat capacity C p ). The term C p ΔT/ΔH $ _{v\textit{ap}} $ is known as the Spalding heat transfer number (Abramzon & Sirignano Reference Abramzon and Sirignano1989).

If there is a convection in the gas phase that is much stronger than conduction when the Péclet number ${\textit{Pe}}= \textit{UR}/\alpha\gg1$ , then h can be enhanced by the flow speed U according to

(4.5) \begin{equation} {\textit{Nu}}\sim {\textit{Pe}}^{\phi }, \end{equation}

where φ is the exponent, depending on detailed flow structures. If U is a given imposed flow speed and ΔT is fixed, as in forced convection, then (4.5) implies NuR φ , leading to n = 2 – φ in (4.3). This results in a negative departure from the D 2-law (Fang et al. Reference Fang, Tseng, Yang and Wei2026). For example, under strong forced convection driven by a constant airflow where φ $=$ 1/2, the corresponding exponent is n $=$ 1.5 (Starinskaya et al. Reference Starinskaya, Miskiv, Nazarov, Terekhov, Terekhov, Rybdylova and Sazhin2021).

In droplet combustion under gravity, however, ΔT will not only set up a buoyant convection to affect Nu in (4.5), but also vary with the droplet radius R that determines the fuel consumption and degree of combustion. Such convection is set up by the red-hot flame of width 2W, causing an upward buoyant flow with velocity (Deen Reference Deen1998)

(4.6) \begin{equation} U\sim \left(g\beta {\Delta} T W\right)^{1/2}. \end{equation}

This velocity scale comes from balancing the driving buoyancy force ρβ $\Delta T$ g W 3 to the inertial force ρU 2 W 2 on the fluid within the lower envelope of the flame where burning is most intense. Here, β ( $=$ 1/T ) is the volume expansion coefficient of the air (of density ρ at temperature T $=$ 300 K). Alternatively, (4.6) can also be obtained by balancing the buoyancy term ρ β $\Delta T$ g to the inertial term ρ $\boldsymbol{v}$ ·∇ $\boldsymbol{v}$ ρ U 2/W in the Navier–Stokes equation for the velocity field $\boldsymbol{v}$ of the airflow around the droplet under the Boussinesq approximation (Tritton Reference Tritton1988). Since $\boldsymbol{v}$ is driven by the flame rather than the droplet, the flame width W serves as the more appropriate length scale in this context.

It is important to note that the droplet merely passively receives heat transfer from the up-flowing convective vapour stream resulting from this natural convection. In other words, from the droplet’s standpoint, since the buoyant convection here is not generated directly by the droplet but indirectly by the flame, the heat transfer to the droplet is of forced convection type. This is fundamentally different from the combustion of a solid particle, where heat transfer is driven by natural convection from a flame situated directly adjacent to the particle (Potter & Riley Reference Potter and Riley1980; Wang et al. Reference Wang, Cui, Bi, Liu, Dong, Xing, Li and Liu2020a ). In the latter case, the induced buoyant velocity scale is U∼ ( $\Delta \textit{TR}$ )1/2, and heat transfer is typically characterised by the Grashof number Gr = ΔTR 3/ν 2 that measures the strength of buoyancy-induced flow relative to viscous dissipation, where ν is the kinematic viscosity of the gas phase. Since the flame size is comparable to the particle size 2R, Gr (which is proportional to R 3) may not be very large when the particle is small. As a result, buoyant convection may be relatively weak compared to that in droplet combustion where the flame is significantly greater than the droplet. A similar situation can also occur in droplet evaporation, where the droplet is typically small and buoyant convection is negligible, thereby allowing the classical D 2-law to remain valid.

Returning to our case, although Nu varies with Pe in a manner like (4.5), this relationship can be recast in terms of Gr. Specifically, by substituting the flame-induced buoyancy velocity scale from (4.6) into Pe = UR/α, we obtain Pe = (ΔTR 2 W/α 2)1/2, which can be rewritten in terms of Gr and the Prandtl number Pr = ν/α as Pe = Gr 1/2 Pr (W/R)1/2. Here, the additional factor (W/R)1/2 accounts for the influence of the flame structure on the droplet. This formulation highlights the indirect but important role of Gr in describing the buoyancy-enhanced heat transfer experienced by the droplet arising from the flame.

Since the flame is the place where the convection is generated, it acts like a permeable hot shell that entrains the surrounding air towards the droplet on the bottom portion of the flame where burning is the most vigorous. How far the flame is located with respect to the droplet, which is reflected by W/R, can also influence heat transfer to the droplet – the closer the flame, the greater the heat transfer rate to the droplet, similar to heat transfer by jet impingement (Liu, Lienhard & Lombara Reference Liu, Lienhard and Lombara1991). So we further add this flame size effect to (4.5) as

(4.7) \begin{equation} {\textit{Nu}}\sim {\textit{Pe}}^{\phi }\left(W/R\right)^{-\omega }, \end{equation}

with ω (>0) being the exponent to account for the degree of this flame size effect.

Since both ΔT and W in (4.6) are also part of the results waiting to be determined, how the heat flux term Nu ΔT in (4.3) depends on R for determining the shrinkage exponent n in (1.2) requires the knowledge of how ΔT and W vary with R. Hence additional conditions are needed to relate ΔT and W to each other.

First, we anticipate that W would tend to be suppressed by U. To find their relationship, we consider the lower envelope of the flame into which there is a strong air entrainment due to intense burning. When the vapourised fuel (with diffusivity $\mathscr{D}$ ) diffuses from the droplet (with the vapour phase concentration ρ fuel at the surface) onto the flame, it is quickly consumed by the burning with fresh oxygen at a high supply rate from the air suction below. Hence on the flame, the fuel vapour’s diffusive flux $j_{fuel} \sim \mathscr{D}\rho_{\textit{fuel}}/W$ across the vapour zone has to match the consumed oxygen’s convective flux j O2U Δρ O2 (times the stoichiometry ratio), where Δρ O2 is the consumed oxygen concentration, which is typically very small because oxygen is quite abundant here, i.e. Δρ O2<< ρ fuel. As a result of j fuel j O2, the flame width is suppressed by the flow according to

(4.8) \begin{equation} W\sim \left({\mathscr{D}}/U\right)\left(\rho _{\textit{fuel}}/{\varDelta} \rho _{\mathrm{O2}}\right)\!. \end{equation}

As for the second condition, we consider the fuel mass balance over the entire flame. The fuel vapourised from the droplet has to be consumed by the oxygen supply from the upward airflow due to the burning. This requires the total fuel consumption rate −r rxn V f within the flame (of volume V f and rate of reaction r rxn < 0) to match the total convective flux of consumed oxygen Δρ O2 UW 2 across the flame width:

(4.9) \begin{equation} -r_{\textit{rxn}}V_{f}\sim \rho _{\mathrm{O2}}UW^{2}. \end{equation}

In other words, the rate of fuel consumption can be represented by that of oxygen. For the same reason, the heat generation from the flame can also be expressed in the same manner. This heat provides the heat flow needed to vapourise the droplet, albeit there is some heat loss through the upward airflow on the top of the flame. Therefore, we can relate the heat flow towards the droplet hΔTA = (k/R)Nu ΔTA in (4.3) to the convective heat flow carried by the airflow:

(4.10) \begin{equation} \left(k/R\right)\textit{Nu}{\Delta} T\,A\sim \left(-r_{\textit{rxn}}V_{f}\right)\,|{\Delta} H_{\textit{rxn}}|\sim {\Delta} \rho _{\mathrm{O2}}UW^{2}\,|{\Delta} H_{\textit{rxn}}|, \end{equation}

where $\Delta H$ rxn (<0) is the heat of reaction.

Now having two additional conditions (4.8) and (4.10) to relate $\Delta T$ and W, we can determine how they vary with R. To do so, it is more convenient to scale both with respective inherent scales determined only by fluid properties. For $\Delta T$ , because it is generated from combustion through heat of reaction $\Delta H$ rxn , it is natural to scale it with the temperature equivalent to the heat of burning reaction T rxn = | $\Delta H$ rxn | $ / $ C p for measuring the fuel’s potential to release heat. As for W, it can be measured relative to some inherent length scale $\mathrm{\ell }$ . This length scale can be found by re-expressing the Péclet number Pe = UR/α after substitution of the velocity scale (4.6):

(4.11) \begin{equation} \textit{Pe}\sim \left({\Delta} T/T_{\textit{rxn}}\right)^{1/2}\left(W/\mathrm{\ell }\right)^{1/2}\left(R/\mathrm{\ell }\right), \end{equation}

wherein

(4.12) \begin{equation} \mathrm{\ell }=\big(g\beta T_{\textit{rxn}}/\alpha ^{2}\big)^{-1/3} \end{equation}

represents the inherent buoyancy length for a given fuel–gas system. It is approximately 100 μm at T rxn of typical value 104 K in air with α = 0.2 cm2 s−1. Since the droplets used in our experiments are on a mm scale, they are larger than this inherent length scale, within which the proposed theory is applicable.

Having scaled $\Delta T$ and lengths with T rxn and $\mathrm{\ell }$ , respectively, we can rewrite the energy balance (4.10) in the dimensionless form

(4.13) \begin{equation} \textit{Nu}\left({\Delta} T/T_{\textit{rxn}}\right)\sim f\,\textit{Pe}\left(W/\mathrm{\ell }\right)^{2}\left(R/\mathrm{\ell }\right)^{-2}, \end{equation}

where f = Δρ O2/ρ is the mass fraction of oxygen consumption. With (4.6), W in (4.8) can be related to $\Delta T$ through

(4.14) \begin{equation} W/\mathrm{\ell }\sim \left({\Delta} T/T_{\textit{rxn}}\right)^{-1/3}\big({\textit{Le}}^{-1}\chi \big)^{2/3}, \end{equation}

where χ = ρ fuel /Δρ O2 is the concentration ratio of the vapour fuel to the consumed oxygen, and $Le = {\alpha}/\mathscr{D}$ is the Lewis number. Substituting (4.7) and (4.11) into (4.13) and replacing W in terms of $\Delta T$ using (4.14), we can determine first $\Delta T$ and then W in terms of R as

(4.15a) \begin{align}&\qquad {\Delta} T/T_{\textit{rxn}}\sim {\Lambda} ^{\gamma }\left(R/\mathrm{\ell }\right)^{-\gamma \left(1+\omega +\phi \right)}, \\[-12pt] \nonumber \end{align}
(4.15b) \begin{align}& W/\mathrm{\ell }\sim {\Lambda} ^{-\gamma /3}\left(R/\mathrm{\ell }\right)^{\gamma \left(1+\omega +\phi \right)/3}\big({\textit{Le}}^{-1}\chi \big)^{2/3}, \\[9pt] \nonumber \end{align}

where Λ = f (Le −1 χ)(5 + 2ω− φ)/3 and γ = 3/(4+ω +φ). Equation (4.15a ) indicates that the temperature difference $\Delta T$ – approximately equal to the flame temperature – increases as the droplet shrinks, which is consistent with experimental observations (Shaw & Wei Reference Shaw and Wei2007; Won et al. Reference Verwey and Birouk2018; Meng et al. Reference Meng, Lai, Zhang, Willmott and Zhang2025). This behaviour can be understood by the fact that as the droplet shrinks, more fuel is vapourised and subsequently consumed on the flame, leading to an increase in the flame temperature. Equation (4.15b ) predicts that the flame width W decreases with the droplet size, as a shrinkage droplet releases less vapourised fuel, thereby reducing the amount of combustible material available to sustain the flame. This leads to a narrower flame envelope, consistent with observations during the later stages of combustion.

With (4.14), the Péclet number Pe = UR/α in (4.11) can be written in terms of $\Delta T$ :

(4.16) \begin{equation} \textit{Pe}\sim \left({\Delta} T/T_{\textit{rxn}}\right)^{1/3}\left(R/\mathrm{\ell }\right)\big({\textit{Le}}^{-1}\chi \big)^{1/3}. \end{equation}

With (4.14) again, Nu in (4.7) can also be written in terms of $\Delta T$ :

(4.17) \begin{equation} {\textit{Nu}}\sim \left({\Delta} T/T_{\textit{rxn}}\right)^{\left(\phi +\omega \right)/3}\left(R/\mathrm{\ell }\right)^{\phi +\omega }\big({\textit{Le}}^{-1}\chi \big)^{\big(\phi -2\omega \big)/3}. \end{equation}

Substituting (4.17) and (4.15a ) into the alternative form of (4.3) by writing $\Delta H$ $ _{v\textit{ap}} $ = C p T $ _{v\textit{ap}} $ :

(4.18) \begin{equation} \frac{{\rm d}R^{2}}{{\rm d}t}=-2\alpha \left(\rho /\rho _{L}\right)\textit{Nu}\left(T_{\textit{rxn}}/T_{\textit{vap}}\right) \left({\Delta} T/T_{\textit{rxn}}\right), \end{equation}

we arrive at

(4.19a) \begin{equation} \frac{{\rm d}R^{2}}{{\rm d}t}=-2\alpha \left(\rho /\rho _{L}\right)\left(T_{\textit{rxn}}/T_{\textit{vap}}\right) \big({\textit{Le}}^{-1}\chi \big)^{\left(\phi -2\omega \right)/3}{\Lambda} ^{\left(\phi +\omega +3\right)\gamma /3}\left(R/\mathrm{\ell }\right)^{-m}, \end{equation}

where m is the deviation exponent from the D 2-law, found to be

(4.19b) \begin{equation} m=\frac{3}{4+\omega +\phi }. \end{equation}

After integration of (4.19a ) and writing the result in terms of D, we arrive at a non-square power law given by (1.2) with the shrinkage exponent

(4.20a) \begin{equation} n=2+m. \end{equation}

By integrating (4.19a ) to (1.2) in terms of D, the resulting burning rate constant is

(4.20b) \begin{equation} K=8\times2^{m}\left(1+m/2\right)c\mathrm{\ell }^{2+m}\left(\frac{\alpha }{\mathrm{\ell }^{2}}\right)\left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right)\big({\textit{Le}}^{-1}\chi \big)^{\left(2\phi +\omega +5\right)m/3}f^{\left(\phi +\omega +3\right)m/3}, \end{equation}

with c being a dimensionless numerical pre-factor.

As indicated by (4.20b ), K scales as $\mathrm{\ell }$ 2+m (α/ $\mathrm{\ell }$ 2) and has units mm2+m s−1, consistent with the dimensional requirement of the non-square law D 2+m = $D_{0}^{2+m}$ Kt from (1.2). Unlike the classical K in the D 2-law, which has units mm2 s−1, this rate constant incorporates buoyant convection effects that lead to m>0, reflecting the non-quadratic nature of the law. This dimensional change is essential, as gravitational acceleration g must enter the shrinkage kinetics through the flame-induced buoyancy velocity scale (4.6), which in turn alters the shrinkage exponent n and thereby the units of K.

In other words, rather than introducing buoyant convection as a minor correction to the K in the D 2-law, as done in the study by Law & Williams (1972), the present work shows that buoyant convection fundamentally changes the burning behaviour. It leads to a new shrinkage kinetic law characterised by a different exponent n and a redefined rate constant K. This distinction is crucial, as it marks a shift from additive corrections to a structurally different formulation that captures the dominant role of buoyancy in shaping the burning dynamics of droplets under normal gravity.

It is worth noting that significant efforts have previously been made to develop theories of droplet vapourisation in both normal and burning regimes, most notably by Sirignano and Law. The theoretical framework presented in this study is fundamentally different from theirs.

Sirignano’s theory is mainly focused on spray combustion in which droplet burning occurs while the droplet is moving within a convective flow (Sirignano Reference Sirignano1983). In this case, the heat and mass transfers are of forced-convection type, with the transfer coefficients described by Ranz–Marshall-type correlations based on the classical Spalding model (Abramzon & Sirignano Reference Abramzon and Sirignano1989). This leads to a D 3/2-law (Prakash & Sirignano Reference Prakash and Sirignano1980), which does not adequately capture the droplet vapourisation dynamics observed under buoyant convection. In contrast, our theory is fundamentally different: it explicitly considers buoyant convection by fully incorporating the coupled effects of momentum, heat and mass transfer. This results in a new Dn -law, with n significantly greater than 2, enabling us to capture the enhanced vapourisation behaviour driven by buoyancy (see § 4.2).

Another aspect of Sirignano’s theory is that the shrinkage kinetics can be further influenced by internal conduction and convective effects within the droplet (Sirignano Reference Sirignano1983). However, our results suggest that such influence may not be as significant as previously assumed. In our analysis, the droplet interior is treated as rapid mixed (i.e. infinitely conductive), thereby neglecting internal thermal gradients. Despite this simplification, our model successfully captures the experimentally measured n values across a range of fuels (see § 4.2). This consistency indicates that under the conditions studied, the shrinkage kinetics is primarily governed by external transport mechanisms, and that internal conduction and convection play a secondary role.

In fact, internal vortical fluid motion within the droplet exhibits closed streamlines, which gives rise to very distinctive heat transfer characteristics. Acrivos and his co-workers have shown both theoretically and experimentally that for heat transfer within closed streamlines, Nu approaches a constant in the Pe → ∞ limit – a value several times higher than that in the pure conduction case (Pe = 0) (Pan & Acrivos Reference Pan and Acrivos1968; Robertson & Acrivos Reference Robertson and Acrivos1970; Acrivos Reference Acrivos1971). Unlike the typical power-law relationship Nu∼Pe φ described by (4.5), this behaviour arises because in the Pe → ∞ limit, the isotherms coincide with the streamlines across which heat transfer can be achieved only by conduction. This implies that internal vortical flow can enhance the droplet’s effective thermal conductivity (Abramzon & Sirignano Reference Abramzon and Sirignano1989). For a non-evaporative droplet, the overall heat transfer can be modelled as a combination of internal and external heat transfers in series (Abramzon & Borde Reference Abramzon and Borde1980). When external convection is strong – i.e. when the external Nusselt number exceeds the internal one – the internal vortex mechanism dominates the overall heat transfer. This internally dominated heat transfer behaviour, characterised by a constant Nusselt number in the high Péclet number regime, has been confirmed by Zhang, Yang & Mao (Reference Zhang, Yang and Mao2012) in their simulation study of a droplet in an external flow field.

However, in our case involving an evaporative droplet, the situation is different because internal and external heat transfers occur simultaneously at the interface, and act in parallel to determine the evaporation rate. As a result, when external convection is strong, the influence of internal vortex effects appears to be minor, as the predictions from the enhanced effective conductivity model arising from such effects do not differ significantly from those of the infinite conductivity model (Abramzon & Sirignano Reference Abramzon and Sirignano1989). Combined with the fact that our experimental results deviate substantially from the D 2-law and closely match theoretical predictions based on the assumption of rapid internal mixing, this strongly supports the validity of the infinite conductivity droplet approximation adopted in our theoretical model.

In fact, such infinite conductivity droplet approximation may be even more applicable in our case, as internal mixing can arise during droplet evaporation due to Marangoni thermal convection induced by non-uniform heating (Mandal & Bakshi Reference Mandal and Bakshi2012; Kazemi et al. Reference Kazemi, Saber, Elliott and Nobes2021), which further reinforces the D 2-law (Mandal & Bakshi Reference Mandal and Bakshi2012). During droplet combustion – where the D 2-law no longer holds – this internal convection can become further intensified and acts in the same direction as the upward buoyant convection, driven by higher temperatures at the droplet’s bottom where the flame burns more intensely. Similar to droplet evaporation under an imposed temperature gradient in a convective environment (Shih & Megaridis Reference Shih and Megaridis1996), such combined convection can further enhance the evaporation rate, thereby causing the droplet’s shrinkage behaviour to deviate even more from the classical D 2-law.

As for Law’s theory, it does account for natural convection effects (Law & Williams 1972), but these effects appear merely as a Grashof number (Gr) based correction to the rate constant K within the framework of the classical D2 -law. Strictly speaking, such a correction is valid only in the limit of small Gr, which contrasts with the actual scenario where Gr is typically large. Our theory is fundamentally different, as it does not treat buoyant convection as a small perturbation to the classical framework or as a variant of the D 2-law. Instead, it fully incorporates buoyancy-driven flow into the governing momentum, heat and mass transfer processes, leading to a new Dn -law with n significantly greater than 2. This new formulation captures the dominant role of buoyant convection in shaping the droplet burning behaviour under normal gravity.

4.2. Influence of thermal boundary layer structures on shrinkage kinetics

The above clearly shows that droplet burning under gravity can alter the D 2-law to a non-square law due to buoyant convection, as also displayed in figure 6 for the measured values of n from experiments. While the measured n values are found to span a narrow range, small but systematic variations can still be observed. Such variations are not unexpected as they can arise from different reaction pathways or thermochemistries, depending on chemical nature of fuels (Curran Reference Curran2019; Yang Reference Yang2021). Such fuel-chemistry-dependent nature can significantly affect flame structures and characteristics (Kohse-Höinghaus Reference Kohse-Höinghaus2021). Since the fuel droplet shrinkage kinetics at large are mainly determined by transport processes to display a nearly universal behaviour, the small variations in the measured n values for a variety of fuels may be explained by different flame and flow structures caused by different combustion chemistries of fuels. In this subsection, we account for such variations in terms of distinct thermal boundary layer structures according to different flame appearances seen in experiments.

We begin with the Nusselt number under the strong convection condition. It can be typically established from Nu ∼ (δ/R)−1 (Deen Reference Deen1998) if the scale of the thermal boundary layer thickness δ can be identified for a given flow field. To do so, instead of using the standard boundary layer theory, we find it is much simpler to make use of δ ∼ (α t)1/2 in transient heat conduction problems to establish the scaling for δ:

(4.21a) \begin{equation} \delta \sim \big({\alpha}\,t_{\textit{flow}}\big)^{1/2}, \end{equation}

with time t in δ ∼ (α t)1/2 being replaced by the flow time t flow based on the flow velocity scale u δ within the boundary layer,

(4.21b) \begin{equation} t_{\textit{flow}}\sim R/u_{\delta }. \end{equation}

Using (4.21a ) together with (4.21b ) in Nu ∼ (δ/R)−1 yields

(4.22) \begin{equation} {\textit{Nu}}\sim \left(\frac{u_{\delta }R}{k}\right)^{1/2}. \end{equation}

This simplified approach provides a more general recipe to determine Nu based on the local Péclet number Pe δ = u δ R/α within the boundary layer without having to involve a formal boundary layer analysis. The advantage of (4.22) is that it offers a unified way to determine Nu for three different flow types (i) slipping flow, (ii) shear flow, and (iii) straining flow commonly seen in practice. In fact, as illustrated in figure 9 as well as seen in the analysis below, these different flow types may happen in different burning situations, allowing us to explain why the measured values of n display small variations between approximately 2.6 and 2.7 shown in figure 6.

Flow type (i) corresponds to the situation where the droplet surface becomes slippery (figure 9 a) due to a possible existence of vapour film (Sirignano Reference Sirignano2014). In this case, u δ U, making (4.22) reduced to the well-known result like the one for bubbles (Deen Reference Deen1998):

(4.23) \begin{equation} {\textit{Nu}}\sim {\textit{Pe}}^{1/2}. \end{equation}

So with ν $=$ 1/2 and ω $=$ 0, (4.15a ) and (4.15b ) become

(4.24a) \begin{align}& {\Delta} T/T_{\textit{rxn}}\sim f^{2/3}\big({\textit{Le}}^{-1}\chi \big) \left(R/\mathrm{\ell }\right)^{-1}, \\[-12pt] \nonumber \end{align}
(4.24b) \begin{align}& W/\mathrm{\ell }\sim f^{-2/9}\big({\textit{Le}}^{-1}\chi \big)^{1/3}\left(R/\mathrm{\ell }\right)^{1/3}. \\[6pt] \nonumber \end{align}

The resulting deviation exponent (4.32b ) is m = 2/3, hence the shrinkage exponent from (4.20a ) is

(4.24c) \begin{equation} n=8/3\approx 2.67. \end{equation}

The burning rate constant from (4.20b ) is then

(4.24d) \begin{equation} K=8\times2^{2/3}(4c/3)\mathrm{\ell }^{8/3}\left(\frac{\alpha }{\mathrm{\ell }^{2}}\right)\left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right)\big({\textit{Le}}^{-1}{\chi} \big)^{4/3}f^{7/9}. \end{equation}

The predicted exponent 2.67 from (4.24c ) well captures the measured values n = 2.66 ± 0.21 (tetradecane) and 2.65 ± 0.17 (dodecanol) for less violent burning like figure 4(b).

Flow type (ii) may take place in the scenario where the droplet surface becomes rigidified due to contamination by soot particles (figure 9 b). It has been observed experimentally that such soot contamination on the surface of a droplet can become quite severe in practically used fuels such as kerosene and diesel (Rasid & Zhang Reference Rasid and Zhang2019). The local flow behaviour around the droplet surface in this case is therefore of shear flow type with u δ /R, rendering (4.22) to be like the equation for solid particles (Deen Reference Deen1998):

(4.25) \begin{equation} {\textit{Nu}}\sim {\textit{Pe}}^{1/3}, \end{equation}

giving φ = 1/3 and ω = 0 in (4.7), and m = 9/13 in (4.19b ). So (4.15a ), (4.15b ), (4.20a ) and (4.20b ) become

(4.26a) \begin{align}&\qquad\qquad\qquad\quad {\Delta} T/T_{\textit{rxn}}\sim f^{9/13}\big({\textit{Le}}^{-1}\chi \big)^{14/13}\left(R/\mathrm{\ell }\right)^{-12/13}, \\[-12pt] \nonumber \end{align}
(4.26b) \begin{align}&\qquad\qquad\qquad\qquad W/\mathrm{\ell }\sim f^{-3/13}\big({\textit{Le}}^{-1}\chi \big)^{4/13}\left(R/\mathrm{\ell }\right)^{4/13}, \\[-12pt] \nonumber \end{align}
(4.26c) \begin{align}&\qquad\qquad\qquad\qquad\qquad\qquad n=35/13\approx 2.69, \\[-12pt] \nonumber \end{align}
(4.26d) \begin{align}& K=8\times2^{9/13}\left(35c/26\right)\mathrm{\ell }^{35/13}\left(\frac{\alpha }{\mathrm{\ell }^{2}}\right)\left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right)\big({\textit{Le}}^{-1}\chi \big)^{17/13}f^{10/13}. \\[6pt] \nonumber \end{align}

The predicted exponent 2.69 from (4.26c ) is slightly higher than 2.67 in (4.24c ), closely capturing the measured values n = 2.73 ± 0.22 (kerosene) and 2.72 ± 0.2 (diesel) for incomplete burning like figure 4(c).

Flow type (iii) may happen to the very violent burning situation for highly volatile fuels such as ethanol and hexane. This is because air entrainment beneath the droplet could become so strong that the uprising flow towards the droplet becomes of straining type impinging the droplet (figure 9 c). Such an impinging flow field around the bottom of the droplet is u δ ∼ (U/W)R, with strain rate ∼U/W, which gives

(4.27) \begin{equation} {\textit{Nu}}\sim {\textit{Pe}}^{1/2}\left(W/R\right)^{-1/2}, \end{equation}

according to (4.22). With $\phi$ $=$ 1/2 and ω $=$ 1/2 in (4.7), and m = 3/5 in (4.19b ), the quantities of interest (4.15a ), (4.15b ), (4.20a ) and (4.20b ) are

(4.28a) \begin{align}&\qquad\qquad\qquad\quad {\Delta} T/T_{\textit{rxn}}\sim f^{3/5}\big({\textit{Le}}^{-1}\chi \big)^{11/10}\left(R/\mathrm{\ell }\right)^{-6/5}, \\[-12pt] \nonumber \end{align}
(4.28b) \begin{align}&\qquad\qquad\qquad\qquad W/\mathrm{\ell }\sim f^{-1/5}\big({\textit{Le}}^{-1}\chi \big)^{3/10}\left(R/\mathrm{\ell }\right)^{2/5}, \\[-12pt] \nonumber \end{align}
(4.28c) \begin{align}&\qquad\qquad\qquad\qquad\qquad\qquad n=13/5=2.6, \\[-12pt] \nonumber \end{align}
(4.28d) \begin{align}& K=8\times2^{3/5}\left(13c/10\right)\mathrm{\ell }^{13/5}\left(\frac{\alpha }{\mathrm{\ell }^{2}}\right)\left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right)\big({\textit{Le}}^{-1}\chi \big)^{13/10}f^{4/5}. \\[6pt] \nonumber \end{align}

The exponent 2.6 from (4.28c ) predicted in this case well captures the measured values n = 2.56 ± 0.16 (ethanol), 2.58 ± 0.07 (hexane), 2.57 ± 0.18 (acetone) and 2.60 ± 0.12 (ethyl acetate) for violent burning like figure 4(a).

As shown above, different flow and boundary layer structures can make $\Delta T$ and W vary with R in different manners and hence yield different values of n. Table 1 summarises the various scaling relationships for $\Delta T$ , W, n and K given above.

Table 1. Summary of distinct scaling relationships for the flame–droplet temperature difference ΔT, the flame width W, and the burning rate constant K at different values of the shrinkage exponent n resulting from different flow and thermal boundary layer structures illustrated in figure 8.

4.3. Burning rate constant

As we can establish the non-square power law (1.2) as above to describe the shrinkage behaviour of a burning droplet under the influence of buoyant convection, the burning rate constant K can be readily obtained from the slope of a plot of (D/D 0) n versus t/D 0 n if n is specified. Since the measured values of n only span a narrow range 2.6–2.7, we can use the D 8/3-law to represent such a nearly universal power-law behaviour. This allows us to make use of the D 8/3t plots to determine the values of K from the slopes for the various fuels that we use in this study, as illustrated in figure 10 for representative fuels: ethanol, tetradecane and kerosene. Figure 11 plots the measured values of K against theory’s prediction

(4.29) \begin{equation} K\sim {\alpha}\mathrm{\ell }^{2/3}\left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right), \end{equation}

according to (4.24d ). We find that the measured K values for pure liquid fuels approximately follow (4.29), which supports our theory. The slope of the best-fitted line is approximately 9, which gives the value of the numerical pre-factor c ´ for the scaling expression (4.29). However, the measured K values for practically used fuels such as kerosene, diesel and petrol are significantly lower than those for pure fuels, implying that the corresponding values of the pre-factor c ´ are substantially smaller. If these practically used fuels still obey (4.29), then their much lower values of c ´ compared to those for pure fuels can only be explained by mixture effects in these fuels, rendering c ´ in this case to be composition-dependent. Perhaps because these fuels are originally distilled from crude oils and are far from pure, the degree of incomplete burning could become much more severe than that of pure liquid fuels, with a lot more soot particles being generated. This not only makes their flames become much extended compared to those of pure fuels, but also significantly hinder heat release from the flames to the droplets, explaining why the observed K values appear lower than for pure fuels.

Figure 10. Plots of (D/D 0)8/3 versus t/D 0 8/3 for determining the values of the burning rate constant K for fuel droplet combustion of (a) ethanol, (b) tetradecane and (c) kerosene, corresponding to the combustion sequences shown in figure 4.

Having a close look at how the measured K values distribute for pure fuels in figure 11, we find that they seem to approximately vary according to fuel types – the K values for alcohol, ketone and ester appear to be lower, whereas those for alkane are higher. For the same family of fuels, the K values do not seem to vary monotonically with carbon content. For alkanes, in particular, the lower carbon content fuel such as hexane can even have a comparable or greater value of K compared to values for higher carbon content, such as decane and tetradecane.

We also find that for a more volatile fuel like ethanol, while it appears to display a higher average flame temperature $\Delta T$ than a less volatile fuel like tetradecane (see figures 4 a,b), the former’s K value is smaller than the latter’s. This may be explained using the alternative form of (4.29) due to the inherent buoyancy length $\ell\propto T_{\textit{rxn}}^{-1/3}$ from (4.12):

(4.30) \begin{equation} K\sim \alpha ^{3}\mathrm{\ell }^{-7/3}g^{-1}\left(\frac{\rho }{\rho _{L}}\right)\left(\beta T_{\textit{vap}}\right)^{-1}\propto \frac{{T_{\textit{rxn}}}^{7/9}}{T_{\textit{vap}}}\propto \frac{\left| {\Delta} H_{\textit{rxn}}\right| ^{7/9}}{{\Delta} H_{\textit{vap}} }. \end{equation}

Equation (4.30) reveals that K grows with heat of reaction $\Delta H$ rxn (<0) as $|$ $\Delta H$ rxn $|$ 7/9. This is because according to (4.3), vapourisation of a fuel droplet is driven by $\Delta T$ due to generation of this heat on the flame; the higher $|$ $\Delta H$ rxn $|$ , the larger K. However, also according to (4.3), the rate of fuel mass vapourisation is diminished by the amount of latent heat $\Delta H$ $ _{v\textit{ap}} $ required to vapourise liquid into vapour, rendering a lower value of K when $\Delta H$ $ _{v\textit{ap}} $ is larger. Since $|$ $\Delta H$ rxn $|$ typically increases with boiling temperature more rapidly than $\Delta H$ $ _{v\textit{ap}} $ , this explains why tetradecane has a greater value of K than ethanol.

For practically used fuels, however, their burning characteristics seem different compared to those of pure fuels. To illustrate this, we take kerosene as an example, and make comparison with tetradecane, whose volatility is the lowest among the pure fuels selected here. Although kerosene’s $|$ $\Delta H$ rxn $|$ is comparable to tetradecane’s, its K value is significantly lower than tetradecane’s. This result may be attributed to formation of soot particles caused by incomplete combustion (Avedisan 2000; Wang Reference Wang2011). This may not only reduce the heat release from the flame, but also cause soot contamination on the droplet surface (Rasid & Zhang Reference Rasid and Zhang2019), thereby further lowering the vapourisation rate and hence the pre-factor c ´ in (4.29). Such soot particles are also responsible for the long and red flames observed in the experiments for this type of fuels (see figure 4 c). Perhaps also because the droplet becomes immobilised due to contamination of these particles on the droplet’s surface, the thermal boundary layer structure becomes of shear flow type with the Nusselt number scaling (4.25) to yield n ≈ 2.69 according to (4.26c ), reasonably capturing the measured n values at approximately 2.73 for kerosene, diesel and petrol.

It is worth mentioning that the burning rate constant (4.29) or (4.30) for droplet combustion under gravity is very different from that without gravity described by the D 2-law. In the latter case, the burning rate constant (4.4b) with $\Delta T$ T rxn is

(4.31) \begin{equation} K\left(g=0\right) \sim \alpha \left(\frac{\rho }{\rho _{L}}\right)\left(\frac{T_{\textit{rxn}}}{T_{\textit{vap}}}\right). \end{equation}

Compared to (4.29), one may simply say that (4.29) can be related to (4.31) as

(4.32) \begin{equation} K\sim K\left(g=0\right)\,\mathrm{\ell }^{2/3}. \end{equation}

However, as implied by $\mathrm{\ell }$ according to (4.12), T rxn from combustion on the flame must work with g to set up buoyancy Δρg ∼ ρg β T rxn to drive upward convection for oxygen supply. To illuminate this important dependence on T rxn under the influence of g, (4.30) is a more appropriate form to represent the burning rate constant.

4.4. Possible sources of experimental variations

As demonstrated above, droplet shrinkage kinetics of various fuels approximately follows a universal law with n = 2.6–2.7, with the observed variations attributed to differences in flow and boundary layer structures. While the differences in the measured n values between fuels are found to be statistically significant, it is worth noting that additional factors may also contribute to these variations.

First, the most prominent factor is thermal radiation, which arises from high flame temperature during combustion. Whether this effect is important can be assessed by comparing the radiative heat flux q rad to the convective heat flux q = h $\Delta T$ from the flame to the droplet in (4.1). The radiative heat flux is evaluated as

(4.33) \begin{equation} q_{rad}=\varepsilon \sigma _{B}\big({T_{f}}^{4}-{T_{s}}^{4}\big)\sim \varepsilon \sigma _{B}{T_{f}}^{4}, \end{equation}

where ε ∼ 1 is the emissivity of the droplet, σB = 5.67 × 10−8 W m−2 K−4 is the Stefan–Boltzmann constant, and T f is the flame temperature, typically much higher than the droplet’s surface temperature T s. The convective heat flux q = h $\Delta T$ can be approximated as h T f and evaluated in terms of the Nusselt number Nu = hR/k:

(4.34) \begin{equation} q\sim \left(k/R\right)\textit{Nu} T_{f}, \end{equation}

where Nu can be estimated as Nu ∼ (UR/α) φ (with $\phi\gt 0$ being the exponent), k ∼ 2.62 × 10−2 W mK-1 is the thermal conductivity of the gas phase, and α ∼ 0.2 cm2 s-1 is the corresponding thermal diffusivity.

With the above, the ratio between (4.33) and (4.34) is

(4.35) \begin{equation} q_{rad}/q\sim \left(\frac{\varepsilon \sigma _{B}T_{f}^{3}}{k/R}\right)\left(\frac{UR}{\alpha }\right)^{-\phi }. \end{equation}

Here, $ \epsilon \sigma_{B}T_{\kern-1pt f}^{3}/(k/R) $ represents the ratio of the radiative heat transfer coefficient ε σ B T f 3 due to thermal radiation to the conductive heat transfer coefficient k/R. Equation (4.35) indicates that the effect of thermal radiation is diminished by convection. With typical values R ∼ 1 mm and U ∼ 1 m/s, the ratio given by (4.35) falls in the range 0.3–0.6 for φ = 1/3–1/2. Therefore, thermal radiation is less significant compared to convection. Moreover, since the flux ratio is q rad /q $ \propto $ R λ with ${\lambda} = 1-{\phi}\gt 0$ according to (4.35), and the above estimate is based on R taken as the initial radius R 0 ∼ 1 mm, the ratio will become even smaller as R decreases during combustion. A similar conclusion can also be drawn for droplet evaporation based on the analogous dependence of q rad / q on R (Abramzon & Sazhin Reference Abramzon and Sazhin2005). This is further supported by model calculations, showing that the difference in vapourisation rates with and without radiation is merely 10 % or less at elevated temperatures (Abramzon & Sazhin Reference Abramzon and Sazhin2006).

Variations can also arise from a range of additional factors. The most common ones include droplet asphericity, fibre heating effects, variations in fluid properties, and the formation of soot particles, all of which can introduce complexities into the combustion process.

Droplet asphericity results from deformation of the droplet by the support fibre. Since this issue can be appropriately addressed using the volume-equivalent approach to determine the effective droplet diameter (see § 2.2) with re-confirmation of the D 2-law, it should not pose a significant problem.

The primary concern regarding the support fibre is actually its potential contribution to additional heating of the droplet. However, since we have used a thinner fibre to conduct the experiment, and found that the measured n values consistently fall within a narrow range regardless of the measurement methods, this issue may not be of great concern. In fact, the fibre heating effect in droplet combustion can be shown to be negligible by evaluating the ratio of the heat flow from the fibre to the droplet, Q fibre = h fibre $\Delta T$ fibre A fibre , to the heat flow from the ambient gas to the droplet, Q = h $\Delta \textit{TA}$ :

(4.36) \begin{equation} \frac{Q_{\textit{fibre}}}{Q}=\frac{h_{\textit{fibre}}{\Delta} T_{\textit{fibre}}}{h{\Delta} T}\left(\frac{A_{\textit{fibre}}}{A}\right), \end{equation}

where h fibre , $\Delta T$ fibre and A fibre denote the heat transfer coefficient, temperature difference and heat transfer area between the fibre and the droplet, respectively. In the fibre heating part, h fibre $=$ k fibre /a fibre since heat transfer primarily occurs through conduction across the fibre, where a fibre is the fibre radius and k fibre is its thermal conductivity. Writing h in terms of the Nusselt number Nu as h $=$ (k/R) Nu ∼ (k/R) (UR/α) φ , with Nu ∼(UR/α) φ and using $\Delta T$ fibre $\Delta T$ and A fibre /A ∼ (a fibre /R)2, (4.36) can be approximated as

(4.37) \begin{equation} \frac{Q_{\textit{fibre}}}{Q}\sim \left(\frac{k_{\textit{fibre}}}{k}\right)\left(\frac{a_{\textit{fibre}}}{R}\right) \left(\frac{UR}{\alpha }\right)^{-\phi }. \end{equation}

With k fibre /k∼5, a fibre /R ∼ 10−1 and UR/α ∼50 under U ∼ 1 m s-1, the resulting ratio is 0.07–0.14 for φ = 1/3–1/2. This indicates that the fibre heating effect contributes approximately 10 % of the ambient gas heating, which supports the observation that the measured n values are largely unaffected by the presence of the fibre (see figures 6 a,b). Comparing results for fibre diameters 35 μm and 100 μm (see figure 7 a,b) shows that while the measured n values for 2a fibre = 100 μm are slightly lower, they exhibit a slight increase with boiling point. The trend also persists for 2a fibre = 35 μm, indicating that it can be attributed not to the fibre but to the intrinsic burning characteristics of the fuel.

It is worth noting that compared to droplet combustion, fibre effects in droplet evaporation may become more pronounced. It can be seen from figure 2 that the cross-fibre method (figure 2 b) yields an n value significantly greater than 2, in contrast to the suspended fibre case (figure 2 a), resulting in a more substantial deviation from the D 2-law. The greater sensitivity of droplet evaporation to fibre effects arises from the absence of convection to enhance the heat flow Q from the ambient gas to the droplet – convection would otherwise diminish the heat flow Q fibre from the fibre. Since both Q fibre and Q are mainly through conduction, their ratio becomes

(4.38) \begin{equation} \frac{Q_{\textit{fibre}}}{Q}\sim \left(\frac{k_{\textit{fibre}}}{k}\right)\left(\frac{a_{\textit{fibre}}}{R}\right), \end{equation}

which makes the fibre contribution more significant compared to that in (4.37) in the presence of strong convection. As also indicated by (4.38), the fibre heating effect is important only when the droplet size 2R approaches the fibre diameter 2a fibre near the end of evaporation. This effect is further amplified in droplet evaporation under the cross-fibre method due to increased droplet–fibre contact, but becomes insignificant in droplet combustion due to suppression by strong buoyant convection. The latter also explains why the fibre only has a marginal influence on the measured n values in droplet combustion, regardless of the fibre supporting method, as shown in figure 6.

Regarding variations caused by non-constant fluid properties, it is evident that properties such as density and heat capacity not only change with temperature but also exhibit spatial variations during combustion, which are not accounted for in our theory. Nevertheless, given the fact that the measured n values closely align with theoretical predictions based on the assumption of constant fluid properties, these effects do not qualitatively alter the nature of the phenomenon. This also implies that variations in fluid properties are unlikely to exceed an order of magnitude sufficient to affect the core physics and key characteristics of the phenomenon. Indeed, numerical studies have shown that variations in droplet vapourisation rates due to temperature-dependent fluid properties are typically within 10–25 % (Kassoy & Williams Reference Kassoy and Williams1968; Megaridis Reference Megaridis1993), which falls within the range of experimental uncertainty. These findings suggest that although such effects are present, they do not compromise the validity of the present theoretical framework.

Formation of soot particles is a common occurrence in flame combustion, resulting from incomplete burning that is intrinsically linked to the fuel’s chemical composition and detailed combustion kinetics (Jin et al. Reference Jin, Yuan, Li, Yang, Zhou, Zhao, Li and Qi2023). The narrow range of measured n values across various fuels suggests that soot formation does not significantly affect the droplet shrinkage mechanism, which is mainly transport-controlled, as demonstrated here. Since soot formation is not entirely deterministic but influenced by stochastic events arising from flame instability (Chu et al. Reference Chu, Qi, Feng, Dong, Hong, Qiu and Han2023), the effects inevitably introduce variations in the droplet shrinkage process, resulting in some degree of scatter in the measured data, as observed in figure 5(a ). Since the extent of incomplete burning may be influenced by the initial fuel quantity, which is controlled by the starting droplet diameter D 0, such soot formation effects could cause the coefficient in the expression (4.20b ) for the burning rate constant K to depend on D 0. This might explain the slight variations in the observed shrinkage kinetics with D 0.

5. Concluding remarks

We have demonstrated, both experimentally and theoretically, that the widely used D2 -law does not accurately characterise the shrinkage behaviour of a burning droplet under the influence of gravity. Our study not only uncovers new features that have not been previously reported, but also provides substantial physical insights that are absent from the existing literature.

Experimentally, by taking special care to minimise the effects of the supporting fibre, we have been able to more accurately determine the shrinkage exponent n, revealing a significant deviation from the D2 -law across a variety of fuels. The measured n values consistently fall within a narrow range 2.6–2.7. Despite differences in hydrocarbon composition and combustion chemistry among the fuels, the observation that the measured n values are so similar suggests that the droplet shrinkage kinetics is universal – primarily governed by underlying transport processes that are largely insensitive to fuel-specific chemical and kinetic details.

Nearly the identical degree of departure from the D 2-law across a variety of fuels can be attributed to the prevailing buoyant convection set up by the flame. This insight enables us to develop a new theoretical framework – fundamentally different from existing models that are largely based on the D 2-law – which accurately captures the shrinkage kinetics of a burning droplet, with predicted n values closely matching experimental observations. Unlike conventional natural convection generated by a heated solid, flame-driven buoyant convection produces three distinct effects. First, the upward convective flow draws in ambient air from beneath the droplet, enhancing the oxygen supply to the flame and promoting combustion. Second, this convection significantly increases heat transfer from the flame to the droplet, thereby accelerating vapourisation. Third, it restricts the outward diffusion of vapourised fuel from the droplet to the flame, leading to a narrower flame envelope. These coupled effects of momentum, heat and mass transfer underlie the non-quadratic shrinkage law derived in this work, which successfully captures the experimentally observed values of n across a wide range of fuels.

While the measured values of n fall within a narrow range, small but systematic variations are still observed. These variations are not unexpected, as they may arise from differences in reaction pathways or thermochemical properties that depend strongly on the chemical nature of the fuels (Curran Reference Curran2019; Yang Reference Yang2021). Such fuel-specific combustion chemistry can significantly influence flame structure and behaviour (Kohse-Höinghaus Reference Kohse-Höinghaus2021). Although the overall droplet shrinkage kinetics is largely governed by transport processes – resulting in an approximately universal behaviour – the minor variations in n across different fuels may be attributed to differences in flame and flow structure shaped by the underlying fuel chemistry.

Along this line, our theory predicts three slightly different values of n, namely 2.6, 8/3 ≈ 2.67 and 35/13 ≈ 2.69, well capturing the narrow range n = 2.6–2.7 found in the experiments, as shown in figure 8. These three values of n correspond to three common flow types: straining flow, slipping flow and shear flow. Straining flow yields n = 2.6, which can explain the measured values for highly volatile fuels such as ethanol and hexane whose flames are mostly blue due to violent burning. Slipping flow results in n = 8/3≈2.67, well capturing the measured values for less volatile fuels such as tetradecane and dodecanol. Although the flame in this case is mostly red due to incomplete burning, the droplet’s surface does not seem to be fully contaminated by soot particles and hence remains somewhat slippery, thereby making slipping flow more appropriate to represent the local flow behaviour around the droplet. Shear flow contributes to n = 35/13≈2.69, in close agreement with the measured values for practically used fuels such as kerosene and diesel. This is because in this case, the droplet is seriously contaminated by soot particles due to strongly incomplete burning with a prolonged flame. This makes the droplet’s surface virtually immobilised, rendering the local flow to become of shear flow type. Attributing these small variations in the value of n to distinct thermal boundary layer structures arising from different combustion characteristics reinforces the interpretation that the observed deviations from the classical D 2-law are governed by underlying transport phenomena rather than experimental artefacts. This highlights both the physical significance and predictive capability of the proposed D n -law framework.

Small changes in the value of n across different fuels – arising from distinct burning behaviours – suggest that the degree of deviation from the D2 -law is relatively stable and not strongly influenced by soot contamination. This stability allows us to adopt the D 8/3-law, corresponding to the intermediate value n = 8/3, as a practical and representative model for describing the universal shrinkage kinetics of burning droplets under gravity. Using this law, we further determine the burning rate constant K for each fuel, and find that the experimental values align well with theoretical predictions (see figure 11), providing additional evidence in support of the nearly universal behaviour observed in our experiments.

The reason behind this seemingly universal description for droplet combustion under gravity lies in the existence of the inherent buoyancy length $\mathrm{\ell }$ given by (4.12), which characterises each fuel–gas system. This length scale makes a necessary connection between the characteristic reaction temperature T rxn and the gravitational acceleration g, effectively quantifying the fuel’s burning capability under buoyant conditions. Its influence is particularly evident in the behaviour of the burning rate constant K, which reflects the breakdown of the D 2-law. Specifically, K no longer scales as the thermal diffusivity α of the gas phase – as it does under microgravity or diffusion-limited conditions described by the D 2-law – but instead follows Kα 3 $\mathrm{\ell }$ n−5/g, as a result of buoyant convection driven by gravity.

Further details on how K depends on other factors – such as fuel and oxygen consumption rates – are provided in (4.24d ), (4.26d ) and (4.28d ), each corresponding to different burning scenarios based on the value of n. However, given that the theoretical values of n differ only slightly, and that the experimentally determined K values consistently follow the predicted scaling in (4.29), the D 8/3-law emerges as a practical and reliable relation for describing the shrinking behaviour of a burning droplet under the influence of buoyant convection. Since this law is fundamentally governed by transport phenomena and is relatively insensitive to detailed chemical kinetics, it – and the theoretical framework behind it – provides valuable guidance for predicting and controlling the temporal size evolution of burning fuel droplets. This makes it particularly useful for optimising flame behaviour in various combustion applications based on fuel properties and operating conditions. For example, in spray combustion where forced convection and natural convection co-exist, the present theoretical framework may be extended to incorporate both effects, providing a unified approach for analysing more complex combustion situations.

Figure 11. Plot of the measured values of K versus α $\mathrm{\ell }$ 2/3 (ρ /ρ L ) (T rxn /T $ _{v\textit{ap}} $ ) according to the D 8/3-law. The line is the best fit of the data for a variety of pure liquid fuels, confirming (4.29) predicted by theory.

Acknowledgements

We acknowledge support from the National Science and Technology Council of Taiwan. H.H.W. also thanks Y.-H. Tseng and T.-I. Lin for their assistance on the preparation of this paper.

Declaration of interests

The authors report no conflict of interest.

References

Abramzon, B. & Borde, I. 1980 Conjugate unsteady heat transfer from a droplet in creeping flow. AIChE J. 26, 537543.10.1002/aic.690260403CrossRefGoogle Scholar
Abramzon, B. & Sazhin, S. 2005 Droplet vaporization model in the presence of thermal radiation. Intl J. Heat Mass Transfer 48, 18681873.10.1016/j.ijheatmasstransfer.2004.11.017CrossRefGoogle Scholar
Abramzon, B. & Sirignano, W.A. 1989 Droplet vaporization model for spray combustion calculations. Intl J. Heat Mass Transfer 32, 16051618.10.1016/0017-9310(89)90043-4CrossRefGoogle Scholar
Abramzon, B. & Sazhin, S. 2006 Convective vaporization of a fuel droplet with thermal radiation absorption. Fuel 85, 3246.10.1016/j.fuel.2005.02.027CrossRefGoogle Scholar
Acrivos, A. 1971 Heat transfer at high Péclet number from a small sphere freely rotating in a simple shear field. J. Fluid Mech. 46, 233240.10.1017/S0022112071000508CrossRefGoogle Scholar
Aggarwal, S.K., Tong, A.Y. & Sirignano, W.A. 1984 A comparison of vaporization models in spray calculations. AIAA J. 22, 14481457.10.2514/3.8802CrossRefGoogle Scholar
Avedisian, C.T. 2000 Recent advances in soot formation from spherical droplet flames at atmospheric pressure. J. Propuls. Power 16, 628635.10.2514/2.5619CrossRefGoogle Scholar
Chauveau, C., Birouk, M., Halter, F. & Gökalp, I. 2019 An analysis of the droplet support fiber effect on the evaporation process. Intl J. Heat Mass Transfer 128, 885891.10.1016/j.ijheatmasstransfer.2018.09.029CrossRefGoogle Scholar
Chen, U.A., Yang, C.Y., Yang, S.Y. & Wei, H.H. 2024 Defying the D 2-law in fuel droplet combustion under gravity. Phys. Fluids 36, 097171.10.1063/5.0225223CrossRefGoogle Scholar
Chow, M.R., Ooi, J.B., Chee, K.M., Pun, C.H., Tran, M.-V., Leong, J.C.K. & Lim, S. 2021 Effects of ethanol on the evaporation and burning characteristics of palm-oil based biodiesel droplet. J. Energy Inst. 98, 3543.10.1016/j.joei.2021.05.008CrossRefGoogle Scholar
Chu, H., Hong, R., Dong, W., Zhang, H., Ma, X. & Chen, L. 2024 Recent advances in soot formation mechanisms: oxidation and oxidation-induced fragmentation. Fuel 371, 132046.10.1016/j.fuel.2024.132046CrossRefGoogle Scholar
Chu, H., Qi, J., Feng, S., Dong, W., Hong, R., Qiu, B. & Han, W. 2023 Soot formation in high-pressure combustion: status and challenges. Fuel 345, 128236.10.1016/j.fuel.2023.128236CrossRefGoogle Scholar
Curran, H. & J, 2019 Developing detailed chemical kinetic mechanisms for fuel combustion. Proc. Combust. Inst. 37, 5781.10.1016/j.proci.2018.06.054CrossRefGoogle Scholar
Deen, W.M. 1998 Analysis of Transport Phenomena. Oxford University Press.Google Scholar
Faeth, G.M. 1977 Current status of droplet and liquid combustion. Prog. Energy Combust. Sci. 3, 191224.10.1016/0360-1285(77)90012-0CrossRefGoogle Scholar
Dalla Barba, F., Wang, J. & Picano, F. 2021 Revisiting D2-law for the evaporation of dilute droplets. Phys. Fluids 33, 051701.10.1063/5.0051078CrossRefGoogle Scholar
Fang, C.-A., Tseng, Y.-H., Yang, S.-Y. & Wei, H.-H. 2026 Beyond the D2-law: combustion: a unified approach to unveiling shrinkage kinetics across diverse droplet vaporization processes. J. Fluid Mech. (in press).Google Scholar
Ferrão, I.A.S., Mendes, TS.M., Mendes, M.A.A., Moita, A.S.O.H. & Silva, A.R.R. 2025 Influence of aluminum nanoparticles in alternative fuel: single droplet combustion experiments and modeling. Fuel 379, 132850.10.1016/j.fuel.2024.132850CrossRefGoogle Scholar
Godsave, G.A.E. 1953 Studies of the combustion of drops in a fuel spray – the burning of single drops of fuel. Symp. (Int.) Combust. 4, 818830.10.1016/S0082-0784(53)80107-4CrossRefGoogle Scholar
Han, K., Yang, B., Zhao, C., Fu, G., Ma, X. & Song, G. 2016 Experimental study on evaporation characteristics of ethanol–diesel blend fuel droplet. Expl Therm. Fluid Sci. 70, 381388.10.1016/j.expthermflusci.2015.10.001CrossRefGoogle Scholar
Jin, H., Yuan, W., Li, W., Yang, J., Zhou, Z., Zhao, L., Li, Y. & Qi, F. 2023 Combustion chemistry of aromatic hydrocarbons. Prog. Energy Combust. Sci. 96, 101076.10.1016/j.pecs.2023.101076CrossRefGoogle Scholar
Kassoy, D.R. & Williams, F.A. 1968 Variable property effects on liquid droplet combustion. AIAA J. 6, 19611965.10.2514/3.4907CrossRefGoogle Scholar
Kazemi, M.A., Saber, S., Elliott, J.A.W. & Nobes, D.S. 2021 Marangoni convection in an evaporating water droplet. Intl J. Heat Mass Transfer 181, 122042.10.1016/j.ijheatmasstransfer.2021.122042CrossRefGoogle Scholar
Kohse-Höinghaus, K. 2021 Combustion in the future: the importance of chemistry. Proc. Combust. Inst. 38, 156.10.1016/j.proci.2020.06.375CrossRefGoogle Scholar
Lai, Y.W. & Pan, K.L. 2023 Disruption and microexplosion mechanisms of burning alcohol droplets with the addition of nanoparticles. Combust. Flame 256, 112958.10.1016/j.combustflame.2023.112958CrossRefGoogle Scholar
Law, C.K. 1982 Recent advances in droplet vaporization and combustion. Prog. Energy Combust. Sci. 8, 171201.10.1016/0360-1285(82)90011-9CrossRefGoogle Scholar
Law, C.K. & Williams, A. 1972 Kinetics and convection in the combustion of alkane droplets. Combust. Flame 19, 393405.10.1016/0010-2180(72)90009-0CrossRefGoogle Scholar
Li, F., Tian, J., Han, K., Bao, L., Meng, K. & Lin, Q. 2021 Expansion and combustion of droplets that contain long-chain alcohol alternative fuels. Phys. Fluids 33, 017117.10.1063/5.0039182CrossRefGoogle Scholar
Liu, X., Lienhard, J.H. & Lombara, V.J.S. 1991 Convective heat transfer by impingement of circular liquid jets. J. Heat Transfer 113, 571582.10.1115/1.2910604CrossRefGoogle Scholar
Mandal, D.K. & Bakshi, S. 2012 Internal circulation in a single droplet evaporating in a closed chamber. Intl J. Multiphase Flow 42, 4251.10.1016/j.ijmultiphaseflow.2012.01.008CrossRefGoogle Scholar
Megaridis, C.M. 1993 Liquid-phase variable property effects in multicomponent droplet convective evaporation. Combust. Sci. Technol. 92, 291311.10.1080/00102209308907676CrossRefGoogle Scholar
Meng, X., Lai, Y., Zhang, Z., Willmott, J. & Zhang, Y. 2025 Experimental investigation of kerosene single droplet ignition and combustion under simulated high-altitude pressure and temperature conditions. Fuel 401, 135825.10.1016/j.fuel.2025.135825CrossRefGoogle Scholar
Misyura, S.Y. 2018 Effect of various key factors on the law of droplet evaporation on the heated horizontal wall. Chem. Engng Res. Des. 129, 306313.10.1016/j.cherd.2017.11.033CrossRefGoogle Scholar
Murakami, Y., Nomura, H. & Suganuma, Y. 2021 Experimental study on unsteadiness of n-decane single droplet evaporation and effect of natural convection on droplet evaporation at high pressures and temperatures. Trans. JSASS Aero. Tech. Jpn. 19, 647653.Google Scholar
Nomura, H., Ujiie, Y., Rath, H.J., Sato, J. & Kono, M. 1996 Experimental study on high-pressure droplet evaporation using microgravity conditions. Symp. (Int.) Combust. 26, 12671273.10.1016/S0082-0784(96)80344-4CrossRefGoogle Scholar
Pan, Y.-F. & Acrivos, A. 1968 Heat transfer at high Péclet number in regions of closed streamlines. Intl J. Heat Mass Transfer 11, 439444.Google Scholar
Parveg, A.S.M.S. & Ratner, A. 2025 A comprehensive review of liquid fuel droplet evaporation and combustion behavior with carbon-based nanoparticles. Prog. Energy Combust. Sci. 106, 101198.10.1016/j.pecs.2024.101198CrossRefGoogle Scholar
Prakash, S. & Sirignano, W.A. 1980 Theory of convective droplet vaporization with unsteady heat transfer in the circulating liquid phase. Intl J. Heat Mass Transfer 23, 253268.10.1016/0017-9310(80)90113-1CrossRefGoogle Scholar
Potter, J.M. & Riley, N. 1980 Free convection from a heated sphere at large Grashof number. J. Fluid Mech. 100, 769783.10.1017/S0022112080001395CrossRefGoogle Scholar
Priyadarshini, S., Kushari, A. & Rao, D.C.K. 2024 Atomization behavior of burning stable emulsion droplets. Fuel 371, 131989.10.1016/j.fuel.2024.131989CrossRefGoogle Scholar
Ranz, W.E. & Marshall, W.R.J. 1952 Evaporation from drops. Chem. Eng. Prog. 48, 141146.Google Scholar
Rasid, A.F.A. & Zhang, Y. 2019 Combustion characteristics and liquid-phase visualization of single isolated diesel droplet with surface contaminated by soot particles. Proc. Combust. Inst. 37, 34013408.10.1016/j.proci.2018.08.023CrossRefGoogle Scholar
Robertson, C.R. & Acrivos, A. 1970 Low Reynolds number shear flow past a rotating circular cylinder. Part 2. Heat transfer. J. Fluid Mech. 40, 705718.10.1017/S0022112070000393CrossRefGoogle Scholar
Sato, J., Tsue, M., Niwa, M. & Kono, M. 1990 Effects of natural convection on high-pressure droplet combustion. Combust. Flame 82, 142150.10.1016/0010-2180(90)90093-7CrossRefGoogle Scholar
Saufi, A.E., Frassoldati, A., Faravelli, T. & Cuoci, A. 2021 Interface-resolved simulation of the evaporation and combustion of a fuel droplet suspended in normal gravity. Fuel 287, 119413.10.1016/j.fuel.2020.119413CrossRefGoogle Scholar
Sazhin, S.S. 2017 Modelling of fuel droplet heating and evaporation: recent results and unsolved problems. Fuel 196, 69101.10.1016/j.fuel.2017.01.048CrossRefGoogle Scholar
Shang, W., Yang, S., Xuan, T., He, Z. & Cao, J. 2020 Experimental studies on combustion and microexplosion characteristics of n-alkane droplets. Energy Fuels 34, 1661316623.10.1021/acs.energyfuels.0c02904CrossRefGoogle Scholar
Shaw, B.D. & Wei, J.B. 2007 Propanol droplet flammability and combustion in air-diluent environments under normal and reduced gravity. Combust. Sci. Technol. 179, 12051223.10.1080/00102200601057394CrossRefGoogle Scholar
Shih, A.T. & Megaridis, C.M. 1996 Thermocapillary flow effects on convective droplet evaporation. Intl J. Heat Mass Transfer 39, 247257.10.1016/0017-9310(95)00137-XCrossRefGoogle Scholar
Sirignano, W. A. 1983 Fuel droplet vaporization and spray combustion theory. Prog. Energy Combust. Sci. 9, 291322.10.1016/0360-1285(83)90011-4CrossRefGoogle Scholar
Sirignano, W.A. 2014 Advances in droplet array combustion theory and modeling. Prog. Energy Combust. Sci. 42, 5486.10.1016/j.pecs.2014.01.002CrossRefGoogle Scholar
Spalding, D.B. 1953 The combustion of liquid fuels. Symp. (Int.) Combust. 4, 847864.10.1016/S0082-0784(53)80110-4CrossRefGoogle Scholar
Starinskaya, E.M., Miskiv, N.B., Nazarov, A.D., Terekhov, V.V., Terekhov, V.I., Rybdylova, O. & Sazhin, S.S. 2021 Evaporation of water/ethanol droplets in an air flow: experimental study and modeling. Intl J. Heat Mass Transfer 177, 121502.10.1016/j.ijheatmasstransfer.2021.121502CrossRefGoogle Scholar
Téré, D., Christian, T.G., Kayaba, H., Boubou, B., Sayouba, S., Tizane, D., Jean, K., Oumar, S., Belkacem, Z. & Antoine, B. 2022 Evaporation and combustion of a drop of liquid fuel – a review. Smart Grid Renew. Energy 13, 2854.10.4236/sgre.2022.132003CrossRefGoogle Scholar
Tritton, D.J. 1988 Physical Fluid Dynamics, 2nd edn. Oxford University Press.Google Scholar
Verwey, C. & Birouk, M. 2018 Experimental investigation of the effect of natural convection on the evaporation characteristics of small fuel droplets at moderately elevated temperature and pressure. Intl J. Heat Mass Transfer 118, 10461055.10.1016/j.ijheatmasstransfer.2017.11.038CrossRefGoogle Scholar
Wang, H. 2011 Formation of nascent soot and other condensed-phase materials in flames. Proc. Combust. Inst. 33, 4167.10.1016/j.proci.2010.09.009CrossRefGoogle Scholar
Wang, S., Cui, G., Bi, H., Liu, C., Dong, Z., Xing, X., Li, Z. & Liu, J. 2020 a Effect analysis on flame characteristics in the combustion of methane hydrate spheres under natural convective flow conditions. J. Nat. Gas Sci. Engng 83, 103578.10.1016/j.jngse.2020.103578CrossRefGoogle Scholar
Wang, J., Huang, X., Qiao, X., Ju, D. & Sun, C. 2020 b Experimental study on effect of support fiber on fuel droplet vaporization at high temperatures. Fuel 268, 117407.10.1016/j.fuel.2020.117407CrossRefGoogle Scholar
Wang, Z., Yuan, B., Huang, Y., Cao, J., Yuzhou Wang, Y. & Cheng, X. 2022 Progress in experimental investigations on evaporation characteristics of a fuel droplet. Fuel Process. Technol. 231, 107243.10.1016/j.fuproc.2022.107243CrossRefGoogle Scholar
Won, J., Baek, S.W. & Kim, H. 2018 Autoignition and combustion behavior of emulsion droplet under elevated temperature and pressure conditions. Energy 163, 800810.10.1016/j.energy.2018.08.185CrossRefGoogle Scholar
Yadev, A.K., Chowdhury, A. & Srivastava, A. 2017 Interferometric investigation of methanol droplet combustion in varying oxygen environments under normal gravity. Intl J. Heat Mass Transfer 111, 871883.10.1016/j.ijheatmasstransfer.2017.03.125CrossRefGoogle Scholar
Yang, B. 2021 Towards predictive combustion kinetic models: progress in model analysis and informative experiments. Proc. Combust. Inst. 38, 199222.10.1016/j.proci.2020.11.002CrossRefGoogle Scholar
Yuge, T. 1960 Experiments on heat transfer from spheres including combined natural and forced convection. J. Heat Transfer 82, 214220.10.1115/1.3679912CrossRefGoogle Scholar
Zhang, J., Yang, C. & Mao, Z.-S. 2012 Mass and heat transfer from or to a single sphere in simple extensional creeping flow. AIChE J. 58, 32143223.10.1002/aic.12811CrossRefGoogle Scholar
Figure 0

Figure 1. Illustration of the D2t plot with different values of the shrinkage exponent n in (a), showing that the profiles with n slightly different from 2 can still appear, resembling the classical D2-law. For n< 2 (e.g. n = 1.5), the curve becomes slightly concave, while for n> 2 (e.g. n = 2.5), it exhibits a mildly convex shape. Such convex profiles can occur in droplet combustion under gravity, as shown in (b) with experimental data for tetradecane, ethanol and kerosene in the present work.

Figure 1

Figure 2. Plots of (1 – t/tlife) versus D/D0 for measuring the shrinkage exponent n during ethanol droplet evaporation (based on 10 realisations), using (a) the suspended method (fibre diameter ∼ 35 μm) and (b) the cross-fibre method (fibre diameter ∼ 100 μm). The measured n value in (a) is pretty close to the ideal value of 2, indicating that the supporting fibre has minimal influence and the D2-law holds. In contrast, (b) shows a significantly larger n, suggesting that the presence of cross-fibres notably affects the evaporation process.

Figure 2

Figure 3. Schematic diagram of the experimental set-up used in the present droplet combustion study.

Figure 3

Figure 4. Sequential images showing the droplet combustion processes of (a) ethanol, (b) tetradecane and (c) kerosene. The lower images in each set provide zoomed-in views of the droplets located in the lower portions of the corresponding frames.

Figure 4

Figure 5. Plots of (1 − t/tlife) against D/D0 for determining the values of the shrinkage exponent n for (a) droplet combustion and (b) evaporation experiments using ethanol, tetradecane and kerosene shown in figure 4.

Figure 5

Figure 6. Measured values of the shrinkage exponent n plotted against the boiling points of various liquid fuels. Both droplet combustion (50 realisations) and evaporation (10 realisations) experiments were conducted under gravity for each fuel. (a) In droplet combustion using the suspended fibre technique (fibre diameter ∼ 35 μm), the values of n obtained by the best-fit method (solid triangles) and the dynamic slope method (open triangles) are approximately 2.6–2.7, consistent with the theoretical prediction $8/3$. These values are notably higher than those approximately 2 observed in the corresponding evaporation experiments (based on the dynamic slope method). (b) Droplet combustion using the cross-fibre technique (fibre diameter ∼ 100 μm) yields similar n values (based on the dynamic slope method), strongly suggesting that the observed departure from the D2-law arises from combustion effects rather than the presence of the fibres.

Figure 6

Figure 7. Effects of the fibre diameter (df) on the shrinkage exponent n for various liquid fuels based on the suspended fibre method. The results are obtained via the direct data fitting method, with data extracted sufficiently away from the fibre to minimise interference. This analysis is to reveal systematic variations of n relative to the theoretical value n = 8/3 predicted by Chen et al. (2024), particularly for low-boiling-point fuels (e.g. hexane and ethanol) and practically used fuels (e.g. diesel and kerosene).

Figure 7

Figure 8. Measured n values may show slight deviations from the theoretical value n = 8/3 predicted by Chen et al. (2024), depending on fuel volatility reflected by boiling point. (a) For more volatile fuels such as ethanol and hexane, their n values are slightly lower than n = 8/3. (b) For less volatile fuels such as tetradecane (C14H30) and dodecanol (C12H25OH), their n values align closely with n = 8/3. (c) For practically used fuels such as diesel and kerosene, their n values become even larger.

Figure 8

Figure 9. Schematic figures for different local flow and boundary layer structures around a burning droplet to account for variations of n observed in the experiments. (a) Slipping flow with little soot particle contamination on the droplet surface, giving n = 8/3 ≈ 2.67. (b) Shear flow with severe soot particle contamination on the droplet surface, which results in n = 35/13 ≈ 2.69. (c) Straining flow resulting from strong airflow impingent arising from vigorous flame burning, yielding n = 2.6.

Figure 9

Table 1. Summary of distinct scaling relationships for the flame–droplet temperature difference ΔT, the flame width W, and the burning rate constant K at different values of the shrinkage exponent n resulting from different flow and thermal boundary layer structures illustrated in figure 8.

Figure 10

Figure 10. Plots of (D/D0)8/3 versus t/D08/3 for determining the values of the burning rate constant K for fuel droplet combustion of (a) ethanol, (b) tetradecane and (c) kerosene, corresponding to the combustion sequences shown in figure 4.

Figure 11

Figure 11. Plot of the measured values of K versus α$\mathrm{\ell }$2/3 (ρ /ρL ) (Trxn/T$ _{v\textit{ap}} $) according to the D8/3-law. The line is the best fit of the data for a variety of pure liquid fuels, confirming (4.29) predicted by theory.