Open Access Article
James H.
Zhang
,
Rohith
Mittapally
,
Abimbola
Oluwade
and
Gang
Chen
*
Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139, USA. E-mail: gchen2@mit.edu
First published on 24th April 2025
Evaporation fluxes from porous evaporators under sunlight have been reported to exceed the solar-thermal limit, determined by relating the incoming solar energy to the latent and sensible heat of water, for applications in desalination and brine pond drying. Although flat two-dimensional (2D) evaporators exceeding the solar limit imply a non-thermal process, tall three-dimensional (3D) solar evaporators can exceed it by absorbing additional environmental heat into its cold sidewalls. Through modeling, we explain the physics and identify the critical heights in which a fin transitions from 2D to 3D evaporation and exceeds the solar-thermal limit. Our analyses illustrate that environmental heat absorption in 3D evaporators is determined by the ambient relative humidity and the airflow velocity. The model is then coarse-grained into a large-scale fin array device on the meters scale to analyze their scalability. We identify that these devices are unlikely to scale favorably in closed environment settings such as solar stills. Our modeling clearly illustrates the benefits and limitations of 3D evaporating arrays and pinpoints design choices in previous works that hinder the device's overall performance. This work illustrates the importance in distinguishing 2D from 3D evaporation for mechanisms underlying interfacial evaporation exceeding the solar-thermal limit.
Broader contextOver the last eight years, many groups have reported that evaporation rates from porous materials under sunlight can exceed the solar-thermal limit calculated by relating the incoming solar energy with the latent and sensible heat of water. There are broadly two classes of materials that are super solar-thermal: 2D and 3D materials. We have shown previously that super solar-thermal 2D evaporators imply a non-thermal evaporation process. In contrast, 3D evaporators can exceed the solar-thermal limit by absorbing environmental energy, which is still a thermal process. Here, we conduct extensive modeling on 3D thermal evaporators to clearly delineate 3D solar interfacial evaporators from 2D evaporators, explain the underlying mechanisms, and evaluate their scalability in large devices for desalination and brine pond drying. Our work pinpoints that 3D evaporators see the most benefit only in open environments under large bulk airflow conditions because environmental heat absorption is governed by the ambient humidity and vapor transport kinetics. This work lays a foundation for understanding the mechanisms underlying the super solar-thermal evaporation limit, and points to directions in exploring their potential applications. |
However, many studies have reported evaporation fluxes that exceed this solar-thermal limit.6,10–19 For 2D solar-driven interfacial evaporating materials in which the solar absorbing area is nominally the same as its evaporating area, reports have demonstrated evaporation fluxes beyond the solar-thermal limit by 2 to 4 times.6,10–13 Our recent work has shown that such high evaporation fluxes imply that water evaporates in the form of clusters, i.e. super solar-thermal, because no region of the evaporation system is below the ambient temperature and the reduced latent heat hypothesis is incorrect.20,21 Although the details of this phenomenon are still under investigation, our group has interpreted such super solar-thermal evaporation as arising from photons directly cleaving off water molecular clusters, which we called the photomolecular effect.10,20,22–24
For tall evaporators, which we call 3D solar interfacial-evaporators and will quantitatively define them later, the evaporating surface area is much larger than its solar absorbing area due to its extended surface similar to fins used in heat transfer devices.15,19,25–29 It has been well appreciated in the field that these structures can exceed the solar-thermal limit due to the structure absorbing additional environmental energy along its sidewalls not exposed to direct sunlight.16 The reported evaporation fluxes normally range between 3 and 5 times the solar-thermal limit27,29 and sometimes even up to 7 times using forced convection,18 based on the projected top cross-sectional area of the absorber. Theoretical studies suggest that water can capillary rise up to meters in height in microporous materials,30 and water has been observed to rise to 100 cm in soil.31 The maximum capillary flux in the 3D evaporator needs to be able to sustain the fins’ total evaporation rate,19 with a past work demonstrating 3D evaporators up to 60 cm tall.17 The high evaporation fluxes have attracted lots of materials development and lab prototype testing. However, very few reports have analyzed the scale-up potential of 3D evaporators beyond a few fins on the decimeter device scale nor have they mechanistically studied the physics of the device.16,18,32–34 Experiments from Chen et al.17 showed that as the array of extended surfaces increase in number, the evaporation flux per structure decreases, illustrating the challenges of scaling up 3D solar evaporators. Yang et al.35 hypothesized that the vapor from the solar absorbing region might re-condense in the evaporative cooling region due to the vapor concentration difference, leading to degraded performance with taller fin heights inside the condenser chamber. Although many works have analyzed 3D evaporators through heat transfer equations,15,16,19,29 very few have explicitly considered vapor transport which drives the environmental heat absorption.18,36 Currently, there lacks a systematic study to help guide these discussions for 3D solar-interfacial evaporators that considers both heat and mass transfer kinetics. No criteria have been established to clearly distinguish 2D from 3D effects, which is important for studying mechanisms behind solar-interfacial evaporation exceeding the solar-thermal limit.
In this work, we will systematically analyze the performance of 3D solar interfacial-evaporators, starting from a single fin and extend to scaled-up systems under both forced and natural convective conditions. First, a simplified model will be constructed to illustrate the performance of a single 3D solar evaporator and reveal the underlying physics. Criteria will be established to demarcate when the evaporating structures can be treated as 3D and when it will reach or exceed the solar-thermal limit. Then, the model will be coarse-grained to a large solar device to study its scale-up potential under forced convective conditions. Our model illustrates that environmental heat input can only occur if the ambient air is below 100% relative humidity and the rate of environmental heat input depends on the airflow velocity and the humidity. Our analysis shows that despite the high performance of a single and a few fins, the performance of large-scale 3D solar evaporator arrays will degrade significantly in low airflow regimes and in closed environments, such as in solar still devices, because of limitations in vapor transport kinetics.
. The foam thickness leads to an additional thickness of the fin tbase not exposed to the ambient air and in contact with a water reservoir underneath at temperature Tbot. The top surface becomes elevated in temperature T and has a high saturated vapor mole fraction cv,s. The sidewalls that are not exposed to direct sunlight drops below the ambient temperature due to evaporative cooling effects, leading to heat absorption from the environment to sustain further evaporation. The solar-thermal limit for 2D evaporation can be calculated by assuming all solar energy is used to heat up and evaporate water, i.e.,![]() | (1) |
We will define the nominal evaporation flux,
, as the total evaporation rate of the fin normalized to only the top projected cross-sectional area Ac. The nominal evaporation flux is the metric commonly reported in previous literature to characterize their performance and governs the solar absorption area when light is incident only on the top surface.
![]() | (2) |
Due to the many variables involved in the model, we have also conducted a sensitivity analysis from the chosen base case by calculating the percentage change in the nominal evaporation flux per percentage change of the variable to understand the most influential terms (Fig. S1, see the ESI†). The most sensitive variables are the fin diameter (−0.93), the ambient temperature (0.67), the fin height (0.64), the solar intensity (0.33), the ambient humidity (−0.31), and the bulk airspeed (0.27). The results are less sensitive to the blackbody emissivity (0.10) and very insensitive to the fin's thermal conductivity (−0.0009). From this analysis, we can appreciate which variables are the most important for single fin performance.
Using the model, we can predict the nominal evaporation flux's dependence on the fin height and diameter (Fig. 1(d)). The variations in the nominal evaporation flux with respect to ambient temperature are shown by the shaded region if the ambient temperature changed by ±10%. Due to the convective heat and mass transfer coefficients’ dependence on both fin diameter and bulk airspeed, we varied the bulk airspeed for the different diameters to maintain a constant sidewall convective heat transfer coefficient of 5 W m−2 K−1 (see the Methods section).
The model predicts that the nominal evaporation flux increases linearly with the fin height due to the increased evaporating surface area assuming no dry out occurs. In a real system, there will be a limit to the height of the fin based on the capillary water pumping in the designed fin. The reason why the nominal evaporation flux increases linearly with fin height is explained with the temperature profiles found previously. As the fin gets taller, it has minimal effects on the temperature profiles near the hot solar-absorbing top and the cooler bottom exchanging heat with the water reservoir below. These features are determined by the balance of the fin's heat conduction along its length with the heat and vapor exchange with the ambient reservoir at the respective ends. Increasing the fin's height only increases the length of the flat temperature profile region in the middle. The energy balance in the flat temperature region is a simple relationship between the evaporative cooling from vapor transport kinetics and the combined convective and radiation heat transfer coefficient with the ambient.
| gm,covCghfg(cv,s(T) − RHcv,s(T∞)) = htot(T∞ − T) | (3) |
From Fig. 1(c), it is evident that the sidewall evaporation flux due to environmental heat flux is about 10 times lower than the maximum solar-thermal rate and the enhanced nominal evaporation flux is due to the cross-sectional area normalization. The wet-bulb temperature is determined by the ambient temperature and relative humidity and it is the lowest temperature the surface can achieve due to evaporative cooling (Fig. S2a, see the ESI†). Using this, we can estimate what the convective sidewall heat transfer coefficient needs to be for the fin's sidewalls to evaporate at the solar-thermal limit due to environmental heat input (Fig. S2b, see the ESI†). It can be seen that except for very high humidity values above 90%, the required sidewall convective heat transfer coefficient is on the order of 100 W m−2 K−1. The radiation heat flux contribution was neglected because its linearized heat transfer coefficient will only be about 5 W m−2 K−1, which is much smaller than the findings in Fig. S2b (ESI†). For a fin diameter of 2.5 cm, this would require an extremely high airspeed at around 17 m s−1.
The performance of 3D evaporators and its environmental heat absorption is governed by the ambient humidity near the fin and the external airflow velocity. If the ambient is near 100% RH, eqn (3) will simplify to the zero solution and no environmental heat absorption can occur. Using this knowledge, we can estimate that the slope governing the increase in the nominal evaporation flux with the fin height is
![]() | (4) |
Due to the strong dependence of the nominal evaporation flux on the ambient humidity and convective conditions, we further plot the nominal evaporation fluxes as a function of these two variables (Fig. 1(e)). Similar to Fig. 1(d), the effects of ambient temperature variations are included using the shaded region as well. As the sidewall convective heat transfer coefficient, and consequentially the convective mass transfer coefficient, increases from 5 to 15 W m−2 K−1, the nominal evaporation fluxes increase by over 2 times at 10% RH and are above the solar-thermal limit due to environmental heat absorption. However, as the ambient humidity increases to 100% RH, all curves predict that the performance will degrade to below the solar-thermal limit for 2D structures and converge towards similar values. At 100% RH, the 3D evaporator will always behave like a 2D evaporator regardless of its height. The structure will continue to evaporate under this condition, albeit at a very low rate compared to the solar-thermal limit, because the solar absorbing top region will be at an elevated temperature compared to the environment. This will cause the saturated vapor mole fraction at the surface to be higher than the vapor mole fraction at 100% RH in the ambient air. In this scenario, the fin's sidewall temperature can never drop below the ambient.
It is important to note that there may be some ambiguity in the definition of what constitutes a wetted-surface area in eqn (2). Many previous 3D interfacial evaporators are highly porous with surface roughness that will cause the surface area to be higher than that of a smooth cylinder. Similar to our previous analysis on 2D evaporators, the key comparison is the characteristic size of the surface roughness to the vapor concentration boundary layer thickness.20 Using the studied conditions from Fig. 1, we can estimate that the vapor concentration boundary layer thickness will be on the order of mm to cm.38 If the surface roughness is significantly smaller, such as on the order of microns, the vapor does not have kinetics to escape from the enhanced surface area into the far field because the macroscopic boundary layer is the rate limiting resistance. If the characteristic size is comparable with the boundary layer, then the fin's perimeter, cross-sectional area, and wetted surface area in eqn (2) should account for the enhanced surface area.
![]() | (5) |
From the discussion on 2D to 3D evaporator transition, it will be meaningful to further identify how tall a fin needs to be, Hcr,th, for the nominal evaporation to reach the solar-thermal limit (Fig. 2(b)). Hcr,2D doesn’t determine the critical height needed to exceed
because the fin first needs to absorb sufficient environment heat to compensate for the heat lost from convective and radiative cooling at the top solar absorbing region. Rather than normalizing to the fin parameter β, Hcr,th is normalized to the fin's geometric aspect ratio p/Ac because the flat-temperature middle region that absorbs environmental energy is independent of the fin's properties if dry-out doesn’t occur. As the ambient humidity increases, Hcr,th increases exponentially because the environmental heat absorption and evaporation are determined by the ambient humidity. For the same reason with Hcr,2D, Hcr,th will also diverge as the ambient humidity approaches 100%. Many previous experiments have fins with fin aspect ratios Hp/Ac on the order of 10 and testing in environmental humidity below 50%.15–19,29Fig. 2(b) suggests that these will achieve super solar-thermal evaporation fluxes.
We have also studied the robustness of the critical heights predicted by the model to variations in the base case parameters (Fig. S3, see the ESI†). The curves illustrate the changes in the critical heights predicted if the thermal conductivity kf (Fig. S3a and b, ESI†), fin diameter D (Fig. S3c and d, ESI†), water reservoir convective heat transfer coefficient hbot (Fig. S3e and f, ESI†), and the base thickness tbase (Fig. S3g–h, ESI†) changed by ±50% of the base values used. The predicted heights to reach the solar-thermal limit Hcr,th is very independent of these variables because the rate of environmental heat absorption is governed only by the airside properties, as shown in eqn (3). The predicted height when 2D evaporators transition to 3D evaporators Hcr,2D has a weak dependence on the fin's thermal conductivity and diameter. Hcr,2D is independent of the water reservoir heat transfer coefficient and the base thickness. The weak dependence on the thermal conductivity is due to the nonlinearity in the model introduced by the evaporative cooling from vapor transport. The weak dependence of Hcr,2D on the fin diameter is due to the convective heat and mass transport on the top cross-sectional surface of the fin. Although the sidewall convective heat transfer coefficient is kept the same, the top cross-sectional's convective coefficients change slightly due to the different scaling with the fin diameter, causing Hcr,2D to shift slightly as well.
| Adev = (SlNrows)(StNcols) | (6) |
![]() | (7) |
In many previously reported outdoor devices, the base of the device is a non-evaporating and non-absorbing insulating material such as white foam instead of an interfacial evaporator. We will show later that this configuration will degrade the true performance of the overall device because a significant amount of solar energy will be wasted when it can be used to further evaporate water. Finally, we define a device evaporation flux that is normalized to the device footprint area to give a representative metric for scaled-up devices that consider the solar absorption of the entire structure.
![]() | (8) |
This metric is only accurate either in the condition when L ≫ H or when θ = 0, with the latter representing the best scenario, and hence our discussion leads to an upper limit in evaporation fluxes.
, external bulk airspeed u∞, and reservoir conditions. However, the large mismatch in length scales between the boundary layer thickness and the device size make FEA simulations of an entire fin array very computationally expensive.
Using the single fin model, we developed a coarse-grained model based on a control volume analysis of the airflow (see the Methods section for details). The list of variables included are shown in Table 1 above. In this model, we draw a series of control volume along the y-axis of the array. Each control volume along the y-axis has dimensions St and Sl such that only one fin is inside. We assume that air only flows along the y-axis such that each fin in the same row behaves identically. We will consider a fin with the same characteristics as the single fin analysis. The transverse spacing St is 4D (10 cm) and the longitudinal spacing Sl is 2D (5 cm). This will cause the extended surface to have about 85% more evaporating surface area than the base plate around it in each control volume. The incoming air is initially at 23 °C and 30% RH. For solar absorption calculations, we will assume that the zenith angle is 0° and that the incoming solar intensity is 1000 W m−2, all absorbed where the solar ray lands (on the top of the fin and the base plate around the fin). The total area, Abase, that absorbs sunlight in each control volume would simply be the product of the two array spacings. This should correspond to the maximum evaporate rate since in this case, the side wall is not heated directly by the solar radiation. In addition to the total device evaporation flux
, we will further define a local device evaporation flux to illustrate the performance of each row of fins, denoted by row index i.
![]() | (9) |
![]() | ||
| Fig. 4 Heat and vapor exchange with airflow through the device for large scale systems. (a)–(c) Properties and performance of a device 50 rows (2.5 m) long with an external bulk airspeed at 1 m s−1. Local air (a) relative humidity and (b) temperature for dark and solar conditions as a function of row number. (c) Contributions from the base plate and the fins’ sidewalls to the local device evaporation flux. (d)–(f) Performance of devices with different lengths and subjected to different bulk airspeeds. (d) Average relative humidity of air throughout the entire device. (e) Ratio of total environmental heat exchange to total solar absorption. (f) Total device evaporation fluxes. The fins have a diameter of 2.5 cm, a height of 10 cm, a thermal conductivity of 0.3 W m−1 K−1, a transverse fin spacing of 10 cm, and a longitudinal fin spacing of 5 cm. The incoming air is initially at 23 °C and 30% RH. The variable lists are included in Table 1 above. | ||
The reason for these trends is revealed in Fig. 4(c), illustrating the contribution of the solar absorbing base plate area and the fin's sidewall contributions to the local device evaporation flux per row. For fins under both dark and solar conditions, the rows near the leading edge of the device contribute a significant amount of evaporation due to the low humidity in the air. In fact, their performances are almost identical because their evaporation fluxes are primarily determined by 3D environmental heat absorption effects. This causes the local air temperature to cool down due to the fins’ sidewall environmental heat absorption. For the base plate near the leading edge, it will heat the air due to convective heat loss under solar conditions and cool the air under dark conditions due to an evaporative cooling effect like the fins’ sidewalls. Further downstream inside the device, the humidity will increase and the fins’ sidewall evaporation becomes less effective, leading to the local fin sidewall evaporation fluxes to degrade rapidly. The base plate downstream under solar conditions continues to both evaporate into and heat up the air due to its enhanced temperatures. Under dark conditions, the base area's local device evaporation flux also decreases rapidly downstream. The balance between the fins’ sidewall heat gain and the base area's heat loss coupled to the changing humidity and local evaporation fluxes create the air temperature minimum under solar conditions. In the initial rows of the device, the local device evaporation flux exceeds the solar-thermal limit due to the low humidity. Further downstream as the air saturates, the local device evaporation flux degrades to below the solar-thermal limit.
These calculations illustrate that the common procedure in previous works, in which the base area is covered with a non-evaporating and non-absorbing material such as white foam, will degrade the device performance significantly. The fins will only have a high device evaporation flux if the local air RH is below 100%. In contrast, the base plate will continue to evaporate water because it heats up from solar absorption, leading to higher saturated vapor molar concentration and enhanced evaporation kinetics. Not utilizing the total solar energy on the entire device will lead to large reductions in performances and this behavior is evident if the device performance is normalized to the entire area footprint rather than only the fin's cross-sectional area.
Our single fin analysis shows that environmental heat gain can only occur if the air flow inside the device is below 100% humidity. Fig. 4(d) illustrates the average humidity of the air across the entire device as a function of its total number of rows (length) and external airspeeds under solar conditions. At low airspeeds, the air will rapidly saturate as the device gets wider. At a bulk airspeed of 2.5 m s−1, the average RH is 70% if the device has 50 rows (2.5 m) in length. This scaling relation is because the rate at which air is replaced in the device scales linearly with the bulk airspeed and the rate of evaporation scales sub-linearly with the bulk airspeed due to the Sherwood number relationship (see the Methods section below). From this analysis, we can appreciate that larger devices and greater environmental gain can be achieved only if the external bulk airspeed is high and the ambient air is dry. Fig. 4(e) shows the total environmental heat gain Δ
env relative to the total solar energy
sun the device absorbs. At lower airspeeds and longer device lengths, the device will lose heat to the environment (Δ
env < 0) because the hot base loses heat faster than the sidewalls can absorb heat. This is because the humid air cannot be replaced by dry air fast enough to enhance the fins’ sidewall evaporation. The corresponding device evaporation fluxes are illustrated in Fig. 4(f), which can exceed the solar-thermal limit under certain conditions. For a given bulk airspeed, there is an optimum number of fins and length the device can be scaled to. Due to the diminished performance of the fins as the device gets larger, the device performance will degrade and eventually be below the solar-thermal limit. If the device is too long, water vapor will evaporate from the hot base and recondense on the fins’ sidewalls further downstream and decrease the device performance even further. From this analysis, the airflow velocity and ambient humidity dictate the potential scale-up of 3D evaporating arrays.
The airflow velocities will depend on the geometry and the boundary layer that forms over the entire device. The natural convection velocities for single fins in these geometries are on the order of cm s−1. For scaled-up devices, the airflow velocity will decrease due to the thicker boundary layers it forms (Fig. 5(a)). The fins near the edge of the device will have the highest heat transfer coefficients and airflow velocities due to it being adjacent to the air reservoir. The interior airflow velocities will be much lower, and it will decrease as the device's transverse width and length (x- and y-axis) increase. The sluggish airflows can lead to the hot vapor near the fins solar absorbing top and base to recondense on the colder sidewalls, as hypothesized in a previous experimental study.35 This will strongly limit the scale-up potential of these evaporating fins under natural convective conditions.
![]() | ||
| Fig. 6 Regime map of device performances. Lines represent single fin performances equal to the solar-thermal limit from Fig. 2(b). The extra area pre-factor in the y-axis accounts for the extra solar irradiation on the base of the device. Example forced convection device evaporation fluxes from Fig. 4(f) are reproduced to illustrate fin device aspect ratio's ability to predict super solar-thermal evaporation fluxes. The length of the devices corresponding to the average RH in the device with bulk airflow u∞ = 0.4 m s−1 is reproduced as well on the top x-axis. The distance between each row is 5 cm. The operating regions of conventional solar still devices are outlined by the gray box. Note that the conventional solar still region is not related to the top axis. | ||
The region in which most reported conventional solar still devices operate are shown by the gray box. Due to the device utilizing natural convection with low induced airflow velocities, the heat transfer coefficients are unlikely to exceed 10 W m−2 K−1. The air inside of the device at steady state must be close to the saturated condition because of the working principle of the device. The regime map clearly illustrates that conventional solar still devices can’t utilize environmental heat input due to the lack of net airflow through the device, the high device RH, and the low airflow velocities. The enhanced outdoor solar still performances reported previously are likely due to a combination of changing solar view factors and improper area normalizations.
We have systematically modeled and explained the physics underlying the performance of 3D evaporating structures. First, the commonly reported nominal evaporation flux and the device evaporation fluxes are distinguished from each other due to the different area normalizations. A single fin model was constructed and was successfully able to reproduce the flat temperature profiles observed in experiments as well as the enhanced nominal evaporation fluxes. The model predicts that the nominal evaporation flux scales linearly with the fin height and that the environmental heat input critically relies on the humidity to be below the saturation point. Using the model, we have identified the non-dimensional critical heights in which a fin will first begin to absorb environmental heat and it absorbs enough environmental heat to nominally evaporate at the solar-thermal limit.
The single fin model was extended to a coarse-grained fin array to study the scale-up performance of these devices by coupling with the vapor and heat transport on the air side as well. Using this model, we have highlighted the potential and limitations in increased device performances by adding fin structures in forced convection situations. The model was able to illustrate the difference in performance between solar and dark evaporation, the environmental heat absorption, and the non-monotonic local air temperature profiles. The model pinpointed that the common procedure of using a foam insulating base plate is ill-advised because the base area can contribute a significant amount of evaporation flux due to enhanced kinetics from solar absorption. The model illustrates that the greatest gain in performance can occur if the external airflow velocity is high so that it can enhance the convective mass transfer coefficients and quickly replenish the humid air inside of the device with dry air. The device length can only be extended up to when the air becomes fully saturated with vapor. Afterwards, device performance will degrade significantly because hot vapor from the solar absorbing regions can recondense on the cooler fins’ sidewalls.
We have identified that the enhanced benefits from 3D evaporating fins are unlikely to translate in large devices for natural convection open environment conditions due to the low airflow velocities. We have also illustrated that environmental heat absorption cannot be the mechanism for enhanced performance under natural convection closed environment conditions, such as in solar still devices, because the air internally is close to 100% RH. Instead, these small-scale devices have misleading improved performances due to the changing solar view factors on the fin and improper area normalization. Through FEA simulations, we have further clarified and illustrated the recondensation effect under natural convective conditions due to the hot vapor region that forms from the evaporating tip, blocking off the vapor transport pathway from the sidewalls.
We have compiled these findings into a regime map to illustrate the performances of single fins, example simulated studies of scaled-up forced convection devices in open environments, and the regime in which conventional solar stills operate in. An extra area ratio pre-factor allows us to approximately map back the overall device performance to the predicted performances of a single fin.
We conclude this paper by emphasizing the fundamental difference between 2D and 3D solar interfacial-evaporators with evaporation fluxes surpassing the solar-thermal limit. As shown in this paper, a 3D evaporator can absorb heat from the environment and achieving evaporation fluxes exceeding the solar-thermal limit.15,16 In this case, the evaporation is still thermally driven. No matter if the evaporators are 2D or 3D, if nowhere in the system is below the ambient temperature, it is impossible to exceed solar-thermal limit based on purely thermal processes because all the heat used for evaporation must come from the solar energy. For 2D evaporators, it is difficult to achieve below ambient temperature unless there is a very high air flowrate.20 Thus, most 2D evaporators tested under natural convection conditions with evaporation fluxes exceeding the solar-thermal limit imply a non-thermal evaporation process, such as via the photomolecular effect that we have discussed in several publications.10,22–24 It is important to point out that the solar absorbing top region of the fin could also evaporate through a non-thermal process as well. We hope this work, together with our investigations on 2D evaporators, will provide stimulus for future understanding of the mechanisms of solar interfacial evaporators exceeding the solar-thermal limit, which will enable better applications.
With such approximations, the differential energy balance equation can be written as the sum of conductive heat transfer along the fin and the convective, evaporative, and radiative heat exchange with the ambient reservoir.
![]() | (10) |
![]() | (11) |
![]() | (12) |
| Nutop = 0.664Re1/2DPr1/3 | (13) |
![]() | (14) |
![]() | (15) |
A system of non-linear equations is formed that relates the heat fluxes at each local cross-section of the fin at a given iteration n.
![]() | (16) |
n is the function vector that becomes zero at the equilibrium temperature profile, the
matrix has constant terms that couple linearly with the temperature vector
n, the
n vector has terms that are non-linearly dependent on the temperature, and the
vector holds constant terms. The Jacobian can then be calculated by taking the multivariable derivative of each term in eqn (16).![]() | (17) |
The newly guessed temperature profile can then be calculated using the Jacobian.
![]() | (18) |
and the old temperature
n using a mixing term λ.![]() | (19) |
The model works by doing a control volume analysis of air flowing downstream along the device and coarse graining the total heat and mass transfer exchange between it and the local row. At the inlet of the ith-row pin fin array, we denote the local air temperature Ti and vapor mole fraction cv,i. The total molar flowrate of “internal” air is ṅair = Cgu∞StH. Žukauskas correlation for tube bundles are used to describe heat transfer between the fins and the air.38,41,42
| Nuconv = Pr0.36fn(ReD,umax) | (20) |
![]() | (21) |
![]() | (22) |
![]() | (23) |
As the airflows over each row, the air's temperature and vapor content will change based on the evaporation and heat exchange.
![]() | (24) |
![]() | (25) |
![]() | (26) |
![]() | (27) |
![]() | (28) |
![]() | (29) |
is the gravity vector, and
is the velocity vector. Using the weakly compressible mode, air's density is coupled with its temperature to induce natural convective flows. Only heat conduction is modeled inside the array structure due to the low flowrates of the liquid involved. The material's thermophysical properties are set to a thermal conductivity of 0.3 W m−1 K−1, a specific heat capacity of 2000 J kg−1 K−1, and a density of 1200 kg m−3.20 The fin was set to 10 cm tall, the base plate set to 2 cm in thickness, the width of the fin set to 2 cm, and the gap between the fins is set to 2 cm. The bottom of the baseplate is set to a Robin heat boundary condition through a heat transfer coefficient of 100 W m−2 K−1 interacting with a reservoir temperature of 23 °C. The air surrounding the fin is initially set to 23 °C and 30% RH. The top of the fins and the top of the base plate absorbs 1000 W m−2 to simulate solar absorption with a zenith angle of 0°. At the fin and air interface, a boundary evaporative cooling term is included that couples the vapor transport solved from eqn (29) with the heat equation in eqn (28) using the latent heat of evaporation. The simulation is then forward marched in time to produce the simulation snapshots.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ee01104c |
| This journal is © The Royal Society of Chemistry 2025 |