Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

ll

OPEN ACCESS

Article
Efficient and antifouling interfacial solar
desalination guided by a transient salt
capacitance model
Jincheng Shi, Xiao Luo, Zhejun
Liu, ..., Changying Zhao,
Xiaokun Gu, Hua Bao

hua.bao@sjtu.edu.cn

HIGHLIGHTS
Circuit-inspired salt capacitance
model for efficient interfacial
evaporation

Importance of thin interfacial


water layer thickness to salt
rejection

High evaporation efficiency of


82% with long-term stability

Maximum daily freshwater yield of


6.0 kg/m2 under natural sunlight

Shi et al. report a circuit-inspired salt capacitance model for efficient and
antifouling interfacial solar desalination by stressing the importance of the
interfacial water layer thickness. Guided by the theoretical model, a low-cost and
scalable design is experimentally demonstrated and achieves high solar-to-vapor
efficiency and freshwater productivity.

Shi et al., Cell Reports Physical Science 2,


100330
February 24, 2021 ª 2021 The Author(s).
https://doi.org/10.1016/j.xcrp.2021.100330
ll
OPEN ACCESS

Article
Efficient and antifouling interfacial solar
desalination guided by a transient
salt capacitance model
Jincheng Shi,1 Xiao Luo,2 Zhejun Liu,1 Junjie Fan,1 Zhouyang Luo,1,3,4 Changying Zhao,2 Xiaokun Gu,2
and Hua Bao1,5,*

SUMMARY
Interfacial solar desalination has been considered as a promising
strategy to address the challenge of freshwater scarcity because
of its high solar-to-vapor efficiency. However, the high evaporation
rate would be restrained if salt crystallization occurs on the evapo-
ration interface. Herein, we propose a capacitance model to
describe transient heat transfer and salt rejecting processes during
interfacial evaporation. This model allows us to identify the water
layer thickness as a new design parameter. With the new under-
standing, the downward water channel ratio can be minimized
to reduce heat loss without salt fouling. Experimental demonstra-
tions show the high efficiency and long-term stability of our design
using extremely low-cost and easily accessible materials. The
maximum daily freshwater yield is as high as 6.0 kg/m2 under natural
sunlight. Our findings not only address the challenge of salt fouling
but also provide insights into transient processes during interfacial
evaporation.

INTRODUCTION
Freshwater scarcity has been considered as a grand challenge for humankind’s
future in the 21st century.1–5 Solar desalination is widely believed to be one of the
most possible solutions for the freshwater scarcity because of the abundant seawater
and solar energy on Earth.6–8 In particular, the concept of interfacial solar desalina-
tion introduced in 2014 achieved a high evaporation rate,9 which has greatly pro-
moted the development of this field.

On the basis of the interfacial solar desalination, tremendous efforts have been
devoted to boosting its evaporation efficiency.10–12 Most of them center their
research around materials development and system design, either by boosting 1University of Michigan–Shanghai Jiao Tong
the solar absorptivity13–25 or lowering the heat loss.19,26–34 Although these improve- University Joint Institute, Shanghai Jiao Tong
University, Shanghai 200240, China
ments can enhance the evaporation efficiency greatly, even to above 80%, the 2Instituteof Engineering Thermophysics, School
impact of salt is neglected in these studies, and the seawater is substituted by fresh- of Mechanical Engineering, Shanghai Jiao Tong
water during the experiments. During the long-term solar desalination process, salt University, Shanghai 200240, China
3KeyLaboratory of Solar Energy Utilization and
crystals unavoidably accumulate on the top evaporation interface, which can block
Energy Saving Technology of Zhejiang Province,
the structure and damage the evaporation efficiency dramatically, thus hindering Hangzhou 311121, China
the practical application and further development of interfacial solar desalination. 4Zhejiang
Energy R&D Institute Co., Ltd.,
Hangzhou 311121, China
Several strategies have been proposed to address the salt fouling issue. One 5Lead contact
possible strategy is to separate the solar absorber and the evaporation surface,35,36 *Correspondence: hua.bao@sjtu.edu.cn
which can effectively avoid salt fouling. However, the evaporation efficiencies are https://doi.org/10.1016/j.xcrp.2021.100330

Cell Reports Physical Science 2, 100330, February 24, 2021 ª 2021 The Author(s). 1
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
ll
OPEN ACCESS
Article

relatively low, likely due to downward heat diffusion from the evaporation surface.
The second strategy is by continuously washing the evaporation surface with
seawater.37 However, this requires further energy to continuously pump water, so
that the entire system becomes complicated. In a major advance in 2018, a very
promising strategy reported by Ni et al.38 is rejecting excess salt effectively through
advection and diffusion, as a result of evaporation-induced salinity gradient along
the water channel. However, in that work, the large water channel area (area ratio
20%) adopted will inevitably cause increased heat loss, because the water channel
itself with thermal conductivity around 0.6 W/(m$K) can be an effective thermal con-
duction path. The evaporation efficiency rate is thus only 56% under one sun. In
some other designs, it is found that there is no salt accumulation in saltwater exper-
iments,39,40 which were believed because of a similar diffusion mechanism. There are
several research studies aiming at optimizing the water channel area ratio in order to
balance the trade-off between evaporation efficiency and salt rejecting.41,42 Kuang
et al.41 reported a natural wood with drilled water channels as salt rejecting path-
ways, which can achieve evaporation efficiency rates of 75% with no salt fouling
at the channel ratio of 7% and would cause salt fouling at the channel ratio of
5%. In other research, Xu et al.42 developed a hierarchical design that can achieve
evaporation efficiency rates of 80% with no salt fouling at the channel ratio of
2.5%. Several questions arise from reviewing the previous studies. Why was an
2.5% channel ratio found to effectively reject salt by Xu et al.,42 but 5% was found
to cause salt fouling by Kuang et al.41? Is it possible to further reduce the channel
ratio to enhance the solar evaporation efficiency while keeping the antifouling per-
formance (i.e., avoid salt crystallization on the absorber and structure)? This moti-
vates us to develop a better fundamental understanding of the salt rejecting process
in the interfacial solar evaporation systems.

In this work, we propose a circuit model that can provide new insights into the mech-
anisms of the salt rejecting process during evaporation. With the circuit model, we
identify another key design parameter, the ‘‘salt capacitance,’’ which is associated
with the transient heat transfer and salt charging process. By tuning this parameter,
we theoretically predict that the effective antifouling property and high energy effi-
ciency can both be achieved by using a minimal water channel ratio. With the new
understanding, we further experimentally demonstrate a low-cost and scalable
interfacial evaporation structure that can achieve high efficiency during long-term
operation in the laboratory and outdoors without salt fouling.

RESULTS
Steady-state analysis
In the typical interfacial evaporation scheme (Figure 1A), the top layer is a solar-
absorbing layer, which has a high solar-to-heat conversion efficiency. Beneath the
top layer is a thermal insulation and water-pumping material. Water pumping can
be achieved either by drilling holes on the thermal insulation material41 or by natural
open pores in a porous media.39 We later call the holes or open pores a water chan-
nel, as shown in Figure 1A. The water channel has multiple effects on the solar desa-
lination system. First, it helps to continuously pump water to the evaporation layer.
Second, it allows salt rejecting by the salt gradient. Third, it causes heat loss because
water generally has much higher thermal conductivity than the insulation materials.
We start the analysis by temporarily assuming that diffusion is the dominating mech-
anism of salt rejecting (excluding advection). According to Fick’s law, the downward
salt diffusion rate per unit evaporation area, J, is proportional to the water channel
ratio, B (ratio of water channel area to the evaporation surface area),

2 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

Figure 1. Typical interfacial solar evaporation design


(A) Schematic of a typical interfacial solar evaporation structure. The top layer is the solar absorbing
layer; beneath the top layer is a thermal insulation and water pumping material.
(B) The downward heat loss and salt diffusion rate as a function of water channel ratio. Downward
heat loss and salt diffusion rate are both proportional to the water channel ratio.

Dc
J = DNaCl Brw ; (Equation 1)
Lt
where DNaCl is the diffusion coefficient of salt in water, rw is the water density, Lt is the
thickness of thermal insulation layer, and Dc is the mass fraction difference of salt to
water between evaporation surface and bulk seawater, Dc = cevap  cN , where cevap
and cN are the mass fraction of salt to water on the evaporation surface and in the
bulk seawater, respectively.

In contrast, by using the parallel heat conduction model, the downward heat loss per
unit evaporation area q is also proportional to the water channel ratio,

DT
q = ðkw B + kt ð1  BÞÞ ; (Equation 2)
Lt
where kw is the thermal conductivity of water, kt is the thermal conductivity of insu-
lation material, and DT is the temperature difference between evaporation surface
and bulk seawater, DT = Tevap  TN , where Tevap and TN are the temperature of
evaporation surface and the bulk seawater, respectively.

From Equations 1 and 2, it is clear that the water channel ratio B is the key factor for
both salt rejecting and heat loss. In our calculation, we assume Lt = 2 cm, cevap = (26/
74)% (assuming saturation), Te = 50 C, and TN = 28 C. The thermophysical proper-
ties are listed in Table S1. Figure 1B shows the heat loss and salt diffusion rate as a
function of B calculated by Equations 1 and 2, from which we find that with the chan-
nel ratio increases, the salt rejecting ability increases while the heat loss also in-
creases. It is thus necessary to fine-tune this parameter to achieve a balance between
salt rejecting and energy efficiency, as has been done by Kuang et al.41

Transient circuit model


The analysis above is reasonable for steady-state conditions, i.e., assuming the sys-
tem is running under continuous long-time solar illumination. However, practical so-
lar desalination is a transient process in nature. The daytime solar illumination is a
function of time, and there is no solar illumination during the nighttime. Under

Cell Reports Physical Science 2, 100330, February 24, 2021 3


ll
OPEN ACCESS
Article

Figure 2. Schematics of antifouling interfacial desalination with salt capacitance design


(A) Schematic of circuit-inspired salt capacitance design. The salt capacitance layer serves as a capacitance for both salt and heat. Solar illumination
serves as a battery, which adds energy to the capacitance layer and deposits salt as a result of evaporation. The insulation material and water channels
beneath work together as a resistance for both salt and energy.
(B) Thermal capacitance and salt capacitance for different salt capacitance layer thicknesses. They are both proportional to the layer thickness.
(C) Temperature and salinity variation over time. When the layer thickness is 0.1 mm, the salt charging time from salinity 3.5 to 26 wt % is about 50 times
larger than the transient heating time (about 2,495 versus 49 s).
(D) Transient heating time and salt charging time for different salt capacitance layer thicknesses. When layer thickness is 1.2 mm, the salt charging time
could reach 7.7 h, and the transient heating time is only 396 s (0.11 h).

illumination, the salinity of the interfacial evaporation layer continuously increases


from 3.5 wt % (the salinity of seawater) because of evaporation. If the salinity reaches
saturation (26 wt %), salt crystals would appear and accumulate on the evaporation
interface, thus hindering further evaporation. During the nighttime, the salinity of the
top layer decreases because of diffusion and advection processes through the water
channel.

This transient ‘‘salt charging’’ process has been largely overlooked in previous
research. The transient process can be analogous to the transient response of a
capacitor (C) through a resistor (R) in an electrical resistor-capacitor (RC) circuit. Fig-
ure 2A shows the typical RC circuit. With a battery, the capacitor C would charge
through the resistor R until it is fully charged. If disconnected from the battery, the
capacitor would discharge itself back through the resistor until the voltage across
the capacitor decays to zero. The transient charging and discharging process can
be tuned by changing the capacitance C.

4 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

Inspired by the transient response features of the RC circuit, we propose a circuit


model that works for both transient salt and heat processes. In contrast with previous
structures, we stress the importance of the thin water layer (‘‘salt capacitance’’) that
usually exists on the top of the insulation material. It serves as a capacitance for both
salt and heat, as shown in Figure 2A. Solar illumination serves as a battery; it adds
energy to the capacitance layer and also deposits salt as a result of evaporation.
The insulation material and water channel beneath work together as a resistance
for both salt and energy. The salt capacitance of the layer per unit evaporation
area Cs can be expressed as:

Cs = rw ðcsat  cN Þ,Ls ; (Equation 3)


where rw is the density of water, csat is the mass fraction of salt to water in saturated
brine, and Ls is the thickness of salt capacitance layer.

Similarly, the heat capacitance of the layer per unit evaporation area Ct can be ex-
pressed as,

Ct = rw cw Tevap  TN ,Ls ; (Equation 4)

where cw is the specific heat capacity of water. Figure 2B shows the capacitance of
salt and heat as a function of the salt capacitance layer thickness. The capacitance
for both salt and heat increases linearly with the layer thickness.

To perform transient circuit analysis, we developed theoretical models based on


mass and energy conservation to demonstrate the transient response of salt and
heat in the capacitance layer (Figure S1; Notes S1 and S2). With the model, we calcu-
late the temperature and salinity of the salt capacitance layer as a function of time by
conservatively assuming no salt downward flow, 90% evaporation efficiency, and 1%
water channel ratio. As shown in Figure 2C, when the layer thickness is 0.1 mm, the
salt charging time from salinity 3.5 wt % to 26 wt % is about 50 times larger than the
transient heating time (about 2,495 versus 49 s), which indicates that the capacitance
layer could largely affect the salt charging time and has little effect on heating time.

The analytical relationship between the salt charging time, ts (h), from salinity 3.5 to
26 wt % and layer thickness, Ls (mm), can also be obtained (Note S1):

ts = 6:4Ls : (Equation 5)
The analytical relationship between the transient heating time, te (h) (defined as the
duration from illumination start until the evaporation surface temperature change
rate smaller than 1 C/min), and layer thickness, Ls (mm), is also given (Note S2),

te = 0:12Ls : (Equation 6)
Based on Equations 3, 4, 5, and 6, it is clear that the thickness of thin capacitance layer is
another key design parameter for interfacial solar evaporation. This parameter has signif-
icant impact on the salt charging process, whereas it has little effect on the thermal pro-
cess. We calculate the salt charging time and transient heating time as a function of Ls
based on Equations 5 and 6, and the results are shown in Figure 2D. From the results,
when layer thickness is 1.2 mm, the salt charging time could reach 7.7 h, and the transient
heating time is only 396 s (0.11 h), which indicates that only a very thin capacitance layer of
1.2 mm could achieve a long salt rejecting time (>7.7 h) and negligible energy efficiency
decrease. Because the total daily solar irradiation is usually less than 8 kWh/m2 in most
places of the world, an 8-h continuous solar irradiation (1 kWh/m2) test is enough, and
the long-time continuous salt rejecting ability is not necessary. We later use 1.2-mm water
layer thickness and 8 kWh/m2 in our indoor experimental demonstration.

Cell Reports Physical Science 2, 100330, February 24, 2021 5


ll
OPEN ACCESS
Article

From the analysis above, we identify the importance of the thin salt capacitance layer
to avoid salt fouling. Another important process, i.e., the salt discharging process,
occurs during both daytime and nighttime. The discharging process involves very
complicated water advection, which is also related to the detailed water channel
design and operation condition. It is left for later discussion in the next section.

Laboratory-scale demonstration
Next, we demonstrate the salt capacitance design using a low-cost system. The
schematic of our design is shown in Figure 3A. The major components of the system
include solar absorber, thermal insulating material, water channel, and the salt
capacitance layer. To make water channels in the insulating material, we drill small
holes in the polystyrene foam (thickness 2 cm, thermal conductivity 0.03 W/
(m,K)) and needle cotton thread in the holes, as shown in Figure 3B. The water chan-
nel ratio is approximately 1% in the structure, estimated from the diameter of the
needle and the number of the holes. To achieve solar absorption and salt capaci-
tance, a gauze dyed in Chinese ink (Figure S2; Note S3) is placed on the top of
the polystyrene foam. The absorptivity of the dyed gauze in wet condition is
measured by a PerkinElmer Lambda 950 spectrometer with an integrating sphere,
and the ultraviolet-visible (UV-vis)-infrared (IR) absorption spectrum is shown in Fig-
ure 3C. More than 98% integrated solar absorption (AM 1.5G) is achieved. The salt
capacitance layer thickness can be controlled by the number of layers of gauze. One
layer of wet gauze is roughly 0.3 mm in thickness as measured by vernier caliper, and
the evaporation surface is a square of 4 cm in length. The evaporation structure is
floating on the simulated bulk seawater (3.5 wt % NaCl) with 4 cm deep in an acrylic
box (the distance between two surfaces is measured as 1.8 cm). It should be noted
that the materials used in our design are extremely low cost. Degreased gauze, ink,
polystyrene foam, and cotton are all easily found in daily life. The entire structure is
very easy to fabricate and can be scalable.

The experimental setup for the measurement of evaporation rate is shown in Fig-
ure 3D. A solar simulator (CEL-S500; Au-light) is used to supply simulated sunlight
(AM 1.5G), and a calibrated solar power meter (CEL-FZ-A; Au-light) is used to cali-
brate the average incoming radiative flux to ensure 1 kW/m2 solar radiation. An
aperture is used to control the size of the light spot.43 The aluminum foil is wrapped
to eliminate the environmental radiation effect (Figure 3E). Thermocouples are used
to continuously record the temperature at different positions in the evaporation
structure. An electronic balance without wind shield is used to continuously record
the mass change of the water.43 The steady-state evaporation efficiency is defined
_ LV =Ein ; where m_ denotes the evaporation rate, hLV is the total enthalpy
as hevap = m,h
of water-to-vapor, including latent heat and sensible heat, and Ein is the incident so-
lar irradiation. The indoor experiments are conducted at a room temperature of
25 C and a humidity of 45% according to the recommendations in Li et al.43

Based on our theoretical analysis, a 1.2-mm water layer is enough to avoid salt
fouling. Therefore, four layers of gauze (1.2 mm total thickness) are used in our
experiment. For comparison, we also use one layer of gauze (0.3 mm) and conduct
the same experiment. Figure 4A shows the appearance of the evaporation surface at
different times after illumination starts. After 8 h of illumination, the evaporation sur-
face of a 1.2-mm layer is free of salt crystals, which verifies the effectiveness of a 1.2-
mm salt capacitance layer in an 8-h antifouling performance. However, for a 0.3-mm
case, the salt fouling becomes visible soon on the evaporation surface (1.5 h), and
the fouling area enlarges over time. This verifies our theoretical analysis, and the salt
capacitance layer thickness can indeed affect salt charging time.

6 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

Figure 3. Evaporation structure with salt capacitance design


(A) Schematic illustration of the salt capacitance design. The major components of the system
include solar absorber, thermal insulating material, water channel, and the salt capacitance layer.
The salt capacitance layer thickness is controlled by the number of gauze layers. Small holes are
drilled in the polystyrene foam for water supply and salt discharging.
(B) Photographs of the bottom, top, and side views of the salt capacitance design. The drilling holes
and hydrophilic cotton thread through the holes (i.e., water channel) provide enough liquid supply
during evaporation. The water channel ratio is 1%. The salt capacitance layer consists of four
layers of gauze (total thickness of 1.2 mm).
(C) Solar absorptivity of the hydrophilic inked degreased gauze.
(D) Schematic of the laboratory-scale experimental setup.
(E) Photograph of the evaporation surface. The aluminum foil is to prevent surrounding light.

As mentioned, whether the salt dissolved in the thin water layer can effectively discharge
is another factor of consideration. We measure the salinity of a 1.2-mm layer over time
during the experiments at 8 and 24 h (0–8 h under 1 kW/m2 illumination and 8–24 h
without illumination). The concentration of salt in the thin layer becomes 24.5 wt % after
8 h. With another 16 h without illumination, it decreases to 7.5 wt %. Based on our salt
rejecting model, and by further considering the downward salt flow, we can simulate the

Cell Reports Physical Science 2, 100330, February 24, 2021 7


ll
OPEN ACCESS
Article

Figure 4. Performance of salt capacitance design for different thicknesses of salt capacitance
layer
(A) Photographs of the evaporation surface under sunlight for 1.2- and 0.3-mm salt capacitance
layers. The evaporation surface of the 1.2-mm layer is free of salt crystals during continuous 8-h
illumination; for 0.3-mm case, the salt fouling becomes visible soon on the evaporation surface, and
the fouling area enlarges over time.
(B) Measured and modeled salinity variation in the 0.3- and 1.2-mm layers over time (8-h sunlight +
16-h dark). Red square is the measuring data of salinity for 1.2-mm salt capacitance layer; red line is
the model result of salinity for 1.2-mm layer; blue square is the measuring data of salinity for 0.3-mm
salt capacitance layer; blue line is the model result of salinity for 0.3-mm layer.
(C) Salt charging time for different thicknesses of salt capacitance layer. Red dot is the measuring
data of salt charging time; black line is the model result of salt charging time.
(D) Transient heating time for different thicknesses of salt capacitance layer. Red dot is the
measuring data of transient heating time; black line is the model result of transient heating time.

continuous variation of salinity over time (Note S1). Figure 4B shows the measured and
modeled salinity variation in the 0.3- and 1.2-mm layers at different times. The salinity of
the 1.2-mm layer could decrease effectively without illumination.

To further explore the transient performance for different layer thicknesses, we mea-
sure the salt charging time and transient heating time, respectively, for different
layer thicknesses. Figure 4C shows the measured salt charging time (red solid circles)
under different layer thicknesses. As expected, the salt charging time almost linearly
increases with water layer thickness. For layer thickness of 1.2 mm, the salt charging
time is 9.52 h. The measurement agrees well with the theoretical calculation using
Equation 5 (solid black line), which verifies the correctness of our theoretical model.
Figure 4D shows the measured transient heating time as a function of the salt capac-
itance layer thickness. Similarly, the transient heating time increases almost linearly
with increasing water layer thickness, and the results agree well with the theoretical
calculation using Equation 6. For layer thickness of 0.3 mm, the minimum transient
heating time is 252 s (0.07 h). The value is only 594 s (0.14 h) for the layer thickness
1.2 mm, which is negligible comparing with the entire solar heating time of 8 h. We
expect that such an increase has little impact on the solar evaporation efficiency.

8 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

Figure 5. Evaporation performance of salt capacitance design under lab conditions


(A) Average steady-state evaporation efficiency and evaporation rate (during the period of no salt
fouling) for different thicknesses of salt capacitance layer. Red square is the evaporation efficiency;
black rhombus is the evaporation rate. The steady-state evaporation efficiency and rate are almost
the same for different layer thicknesses. Error bars indicate one standard deviation.
(B) 10 min average evaporation efficiency during 8-h continuous simulated sunlight for 0.3- and 1.2-
mm salt capacitance layers. Red square is for 1.2-mm salt capacitance layer; blue dot is for 0.3-mm
layer. The evaporation efficiencies are almost identical over time for the 1.2-mm salt capacitance
layer structure; the evaporation efficiency decreases over time for the 0.3-mm layer case. Error bars
indicate one standard deviation.
(C) Evaporation performance of the 1.2-mm salt capacitance layer design. The mass change over
time (red line). Freshwater test (dark dot) is for comparison.
(D) The temperature variation over time in different positions of our structure during evaporation.
The inset shows the different locations: T1 is the evaporation surface; T 2 is the bottom surface of the
thin evaporation layer; T 3 is the upper surface of bulk seawater; T 4 is the bottom surface of the bulk
seawater.

Now that the transient heating time is negligibly increased by the thin salt capacitance
layer, it is critical to verify that the thin salt capacitance layer has no effect on the steady-
state evaporation efficiency. The steady-state evaporation efficiency is measured using
the same lab-scale experiments mentioned above. As Figure 5A shows, the steady-state
evaporation efficiency and rate (before obvious salt fouling appears) are almost the same
for different layer thicknesses, which confirms that the salt capacitance layer has no detri-
mental impact on the steady-state evaporation efficiency. In contrast, the evaporation
efficiency reduces dramatically by the salt fouling. Figure 5B shows the evaporation ef-
ficiency variation during 8-h continuous simulated sunlight (1 kW/m2) for 0.3- and 1.2-
mm layers. The results show that the 10-min average evaporation efficiencies at different
times are almost identical over time for the 1.2-mm salt capacitance layer case. However,
the evaporation efficiency decreases over time for the 0.3-mm layer structure, which is
caused by the salt crystals accumulation on the evaporation surface mentioned above.

To further confirm the evaporation performance of the 1.2-mm salt capacitance


layer, we test the structure in both simulated seawater and freshwater. Figure 5C

Cell Reports Physical Science 2, 100330, February 24, 2021 9


ll
OPEN ACCESS
Article

Figure 6. Long-term performance of 1.2-mm salt capacitance layer under lab conditions
(A) The salinity variation over time in the salt capacitance layer. Red square is the measuring data
using salometer during the experiments; black line is the modeled result (Note S1). The highest and
lowest salinity in the salt capacitance layer gradually stabilize at about 24.5 and 7.5 wt % after
2 days.
(B) Daily average evaporation efficiency (red square) and evaporation rate (black rhombus). The
evaporation efficiency is stable over 5 days, showing an average evaporation efficiency of about
82%. Error bars indicate one standard deviation.

shows the mass change over time of the two cases after the system reaches a quasi-
steady state, which indicates the comparable evaporation rates of the structure in
seawater and freshwater. The evaporation efficiencies are calculated as 82% G
2% according to the sloping of the lines. The corresponding evaporation tempera-
tures are measured as 48 C.

The high evaporation efficiency of our design should be attributed to the small
downward heat loss as a result of the confined water channel (ratio 1%), which is
verified by the temperature recordings at different locations in the structure (Fig-
ure 5D). After 8 h of solar irradiation, the bulk seawater temperature increased by
only 1 C. Although the effect of the water channel ratio on the efficiency has been
discussed in a previous work,41 we verify their conclusion by conducting an experi-
ment with different water channel ratios using our system. As expected, when the
water ratio increases from 1% to 20%, the efficiency continuously decreases from
82% to 58% (Figure S3; Note S4). The detailed energy loss analysis of our struc-
ture can be found in Figure S4 and Note S5.

To demonstrate the long-term stability of the 1.2-mm salt capacitance layer design
in the simulated seawater (3.5 wt % NaCl), we perform a continuous 5-day labora-
tory experiment with 8-h continuous 1 kW/m2 irradiation per day. No salt crystals
are observed in the structure and evaporation surface. Furthermore, the salinity of
the salt capacitance layer is measured over time, as shown in Figure 6A, which is
always below the saturation of 26 wt %. The highest and lowest salinity in the
salt capacitance layer are gradually stable at about 24.5 and 7.5 wt %, respectively,
after 2 days, suggesting the structure can operate with long-term stability. To give
a brief idea of the transient salt response process, we use the theoretical model
mentioned above to simulate the continuous variation of salinity over time (Note
S1), which includes the salt charging process under illumination and the discharg-
ing process under the absence of illumination as a result of the diffusion and
advection along the water channel. The modeled salinity curve is seen in Figure 6A,
showing good agreement with the experimental measurement. Figure 6B shows
the everyday average steady-state evaporation efficiency. The steady-state evapo-
ration efficiency is found to be stable over 5 days, showing an average evaporation
efficiency of about 82%.

10 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

Figure 7. Performance of salt capacitance design in solar still under natural sunlight
(A) Photograph of the outdoor demonstration system. The right is the solar still with salt
capacitance design; the left is the solar still with bottom heating for comparison.
(B) The solar flux on a cloudy day. E solar is the daily solar insolation per unit evaporation area.
(C) The daily water productivity of the two solar stills. Blue bar is for interfacial heating, and orange
bar is for bottom heating.
(D) The daily solar-to-water efficiency of solar still. Blue bar is for interfacial heating, and orange bar
is for bottom heating.

Outdoor demonstration
Following the excellent performance of the laboratory demonstration, outdoor solar
desalination experiments are performed on a roof of Zheneng Chuangye Building
located in Hangzhou from August 16, 2020, to August 20, 2020. As shown in Fig-
ure 7A, the solar still is insulated by extruded polystyrene foam with a thickness of
2 cm (thermal conductivity 0.03 W/(m,K)). The bottom surface of the solar still is
50 3 50 cm. The front and back heights are 10 and 28 cm, respectively. A glass cover
(transmissivity 90%) is used on top of the solar still with inclination angle of about
20 . We place an enlarged evaporation structure as our indoor experiment in the so-
lar still. An identical solar still with bottom heating is used for comparison under the
same condition. Simulated seawater (3.5 wt %) is used in both solar stills. During the
continuous 5 days of outdoor experiment, we observe the evaporation surface hour-
ly from 9:00 a.m. to 5:00 p.m. The condensate water is collected in the measuring
cylinder for 24 h every day at 6:00 a.m. A calibrated solar power meter (CEL-
NP2000-2A; Au-Light) can continuously measure the outdoor solar flux. A typical re-
corded solar flux is shown in Figure 7B.

No salt crystals are observed on the surface of the structure over 5 days, indicating
the antifouling effectiveness of the circuit-inspired salt capacitance design for long-
term outdoor application. Figure 7C shows the water productivity over 5 days, which
indicates the larger water productivity of our design than the bottom heating case.
The maximum water productivity 6.0 kg/(m2,day) of our structure is observed on

Cell Reports Physical Science 2, 100330, February 24, 2021 11


ll
OPEN ACCESS
Article

August 16, while the bottom heating case was 4.7 kg/(m2,day). The water produc-
tivity of 6.0 kg/(m2,day) is higher than most works that have been reported so far
with a single-stage solar still. The solar-to-water efficiency is calculated by hwater =
m,hf =Esolar ; where m denotes the mass of condensate water, hf is the latent heat
of water-to-vapor, and Esolar is the integrated incident solar irradiation. Figure 7D
shows the corresponding solar-to-water efficiency over 5 days, showing a maximum
efficiency 57% of our structure and 45% of the bottom heating case on August 16. It
should be mentioned that the solar stills used in our experiments are relatively sim-
ple, just to verify the effectiveness of our design in outdoor conditions. The solar-to-
water efficiency could be further enhanced by improving the design of solar stills.

DISCUSSION
We underline the transient nature of salt charging and rejecting processes during
interfacial evaporation based on theoretical analysis and experimental demonstra-
tion. Although named as ‘‘interfacial,’’ the evaporation occurs at a finite water layer
thickness. This layer has negligible heat capacitance that has a small effect on the
energy efficiency, whereas its effect is significant on the antifouling performance.
The experimental results show that only a very thin salt capacitance layer
(1.2 mm) would significantly increase the salt charging time, while the negligible
transient heating time increased (9.52 versus 0.14 h). Furthermore, the salt capaci-
tance layer would not decrease the steady-state evaporation efficiency (82%).
Continuous 5-day laboratory experiment results indicate the long-term stability of
our antifouling evaporation structure, which does not need regular cleaning and
maintenance. The outdoor experiment result indicates the effectiveness of the struc-
ture in solar still with solar-to-water efficiency of 57%, with corresponding water pro-
ductivity of 6.0 kg/(m2,day), which is better than the conventional bottom heating
solar still (45%). It should be noted that the materials used in our design are
extremely low cost, about $2.4 per unit evaporation structure (estimation in Note
S6), and the result is also insensitive to the particular materials we choose. These
offer much more flexibility for further development for large-scale and long-term
practical applications.

As we know, the advantage of interfacial heating is to reduce the heat capacity of the
seawater, which leads to a rapid transient heating process and limited heat loss. The
transient process of evaporation and salt accumulation has been largely overlooked
by previous works in the field of interfacial solar desalination. The thickness of the
evaporation layer in previous systems is mostly neglected and depends only on
the thickness of the solar absorber. Based on our understanding, the reason that
some previous designs can achieve antifouling may be the large evaporation layer
thickness. For example, Li et al.39 developed a bilayer melamine foam, which could
achieve salt free under a series of durability tests. It should be caused by the very
large thickness of the melamine foam (15 mm).

In summary, we demonstrated, both theoretically and experimentally, a salt


charging theory and the corresponding evaporation structure based on a thin
salt capacitance layer, which can achieve long-term antifouling performance
without compromising the evaporation efficiency, simplicity, and material cost.
Moreover, we elucidate the role of the transient process in the field of interfacial
solar desalination. We believe the findings in this work not only open up an avenue
for antifouling properties but also provide revitalized attention to the transient
process in this field, thus helping further development of the interfacial solar desa-
lination technology.

12 Cell Reports Physical Science 2, 100330, February 24, 2021


ll
OPEN ACCESS
Article

EXPERIMENTAL PROCEDURES
Resource availability
Lead contact
Further information and requests for resources and reagents should be directed to
and will be fulfilled by the Lead Contact, Hua Bao (hua.bao@sjtu.edu.cn).

Materials availability
This study did not generate new unique reagents.

Data and code availability


The authors declare that data supporting the findings of this study are available
within the paper and the Supplemental information. All other data are available
from the Lead Contact upon request.

Materials
Polystyrene foam was purchased commercially from Basuhe (China). Cotton thread
and gauze were purchased from Xinglong (China). Chinese ink was purchased from
Yidege (China). NaCl was purchased from Mokai (China). All of the reagents were
used without further purification.

SUPPLEMENTAL INFORMATION
Supplemental Information can be found online at https://doi.org/10.1016/j.xcrp.
2021.100330.

ACKNOWLEDGMENTS
We gratefully acknowledge funding support from the Zhejiang Energy R&D Institute
and National Natural Science Foundation of China (No. 51636004) and the support
from University of Michigan-Shanghai Jiao Tong University Joint Institute.

AUTHOR CONTRIBUTIONS
J.S. and H.B. conceived the idea and designed the research project; J.S., X.L., Z.J.L.,
J.F., Z.Y.L., X.G., and H.B. performed the experiments and collected and analyzed
the data; J.S. and H.B. wrote the manuscript; and H.B. supervised the project. All
of the authors discussed the results and commented on the manuscript.

DECLARATION OF INTERESTS
The authors declare no competing interests.

Received: September 22, 2020


Revised: November 21, 2020
Accepted: January 6, 2021
Published: January 26, 2021

REFERENCES
1. Schwarzenbach, R.P., Egli, T., Hofstetter, 3. Mekonnen, M.M., and Hoekstra, A.Y. (2016). 6. Kabeel, A.E., and El-Agouz, S.A. (2011). Review
T.B., Von Gunten, U., and Wehrli, B. Four billion people facing severe water of researches and developments on solar stills.
(2010). Global water pollution and human scarcity. Sci. Adv. 2, e1500323. Desalination 276, 1–12.
health. Annu. Rev. Environ. Resour. 35,
109–136. 4. Oki, T., and Kanae, S. (2006). Global 7. Kabeel, A.E., Muthu Manokar, A.,
hydrological cycles and world water resources. Sathyamurthy, R., Prince Winston, D., El-
2. Hoekstra, A.Y., Mekonnen, M.M., Chapagain, Science 313, 1068–1072. Agouz, S.A., and Chamkha, A.J. (2019). A
A.K., Mathews, R.E., and Richter, B.D. (2012). review on different design modifications
Global monthly water scarcity: blue water 5. Khawaji, A.D., Kutubkhanah, I.K., and Wie, employed in inclined solar still for enhancing
footprints versus blue water availability. PLoS J.-M. (2008). Advances in seawater desalination the productivity. J. Sol. Energy Eng. 141,
ONE 7, e32688. technologies. Desalination 221, 47–69. 031007.

Cell Reports Physical Science 2, 100330, February 24, 2021 13


ll
OPEN ACCESS
Article

8. Sharshir, S.W., Peng, G., Wu, L., Yang, N., Essa, performance solar steam generation. ACS under one-sun irradiation. J. Mater. Chem. A
F.A., Elsheikh, A.H., Mohamed, S.I.T., and Appl. Mater. Interfaces 9, 15052–15057. Mater. Energy Sustain. 5, 17817–17821.
Kabeel, A.E. (2017). Enhancing the solar still
performance using nanofluids and glass cover 21. Wang, Z., Liu, Y., Tao, P., Shen, Q., Yi, N., 32. Liu, H., Zhang, X., Hong, Z., Pu, Z., Yao, Q., Shi,
cooling: Experimental study. Appl. Therm. Zhang, F., Liu, Q., Song, C., Zhang, D., Shang, J., Yang, G., Mi, B., Yang, B., Liu, X., et al. (2017).
Eng. 113, 684–693. W., and Deng, T. (2014). Bio-inspired A bioinspired capillary-driven pump for solar
evaporation through plasmonic film of vapor generation. Nano Energy 42, 115–121.
9. Ghasemi, H., Ni, G., Marconnet, A.M., Loomis, nanoparticles at the air-water interface. Small
J., Yerci, S., Miljkovic, N., and Chen, G. (2014). 10, 3234–3239. 33. Li, Y., Gao, T., Yang, Z., Chaoji, C., Kuang, Y.,
Solar steam generation by heat localization. Song, J., Jia, C., Hitz, E.M., Yang, B., and Hu, L.
Nat. Commun. 5, 4449. 22. Zhou, L., Tan, Y., Wang, J., Xu, W., Yuan, Y., Cai, (2017). Graphene oxide-based evaporator with
W., Zhu, S., and Zhu, J. (2016). 3D self-assembly one-dimensional water transport enabling
10. Tao, P., Ni, G., Song, C., Shang, W., Wu, J., Zhu, of aluminium nanoparticles for plasmon- high-efficiency solar desalination. Nano Energy
J., Chen, G., and Deng, T. (2018). Solar-driven enhanced solar desalination. Nat. Photonics 41, 201–209.
interfacial evaporation. Nat. Energy 3, 1031– 10, 393–398.
1041. 34. Deng, Z., Liu, P.-F., Zhou, J., Miao, L., Peng, Y.,
23. Zhou, L., Tan, Y., Ji, D., Zhu, B., Zhang, P., Xu, J., Su, H., Wang, P., Wang, X., Cao, W., Jiang, F.,
11. Pang, Y., Zhang, J., Ma, R., Qu, Z., Lee, E., and et al. (2018). A novel ink-stained paper for solar
Gan, Q., Yu, Z., and Zhu, J. (2016). Self-
Luo, T. (2020). Solar–thermal water heavy metal treatment and desalination. Solar
assembly of highly efficient, broadband
evaporation: a review. ACS Energy Lett. 5, RRL. 2, 1800073.
plasmonic absorbers for solar steam
437–456.
generation. Sci. Adv. 2, e1501227.
35. Cooper, T.A., Zandavi, S.H., Ni, G.W.,
12. Gao, M., Zhu, L., Peh, C.K., and Ho, G.W. Tsurimaki, Y., Huang, Y., Boriskina, S.V., and
(2019). Solar absorber material and system 24. Zhu, M., Li, Y., Chen, F., Zhu, X., Dai, J., Li, Y.,
Yang, Z., Yan, X., Song, J., Wang, Y., et al. Chen, G. (2018). Contactless steam generation
designs for photothermal water vaporization and superheating under one sun illumination.
towards clean water and energy production. (2018). Plasmonic Wood for High-Efficiency
Solar Steam Generation. Adv. Energy Mater. 8, Nat. Commun. 9, 5086.
Energy Environ. Sci. 12, 841–864.
1701028.
36. Xu, W., Hu, X., Zhuang, S., Wang, Y., Li, X.,
13. Wang, Y., Zhang, L., and Wang, P. (2016). Self- Zhou, L., Zhu, S., and Zhu, J. (2018). Flexible and
floating carbon nanotube membrane on 25. Liu, Y., Chen, J., Guo, D., Cao, M., and Jiang, L.
(2015). Floatable, self-Cleaning, and carbon- salt resistant Janus absorbers by
macroporous silica substrate for highly efficient electrospinning for stable and efficient solar
solar-driven interfacial water evaporation. ACS black-based superhydrophobic gauze for the
solar evaporation enhancement at the air– desalination. Adv. Energy Mater. 8, 1702884.
Sustain. Chem. Eng. 4, 1223–1230.
water interface. ACS Appl. Mater. Interfaces 7,
37. Liu, Z., Wu, B., Zhu, B., Chen, Z., Zhu, M., and
14. Yang, Y., Zhao, R., Zhang, T., Zhao, K., Xiao, P., 13645–13652.
Liu, X. (2019). Continuously producing
Ma, Y., Ajayan, P.M., Shi, G., and Chen, Y.
watersteam and concentrated brine from
(2018). Graphene-based standalone solar 26. Li, X., Xu, W., Tang, M., Zhou, L., Zhu, B., Zhu, S.,
seawater by hanging photothermal fabrics
energy converter for water desalination and and Zhu, J. (2016). Graphene oxide-based
under sunlight. Adv. Funct. Mater. 29, 1905485.
purification. ACS Nano 12, 829–835. efficient and scalable solar desalination under
one sun with a confined 2D water path. Proc. 38. Ni, G., Zandavi, S.H., Javid, S.M., Boriskina,
15. Ito, Y., Tanabe, Y., Han, J., Fujita, T., Tanigaki, Natl. Acad. Sci. USA 113, 13953–13958.
K., and Chen, M. (2015). Multifunctional Porous S.V., Cooper, T.A., and Chen, G. (2018). A salt-
Graphene for High-Efficiency Steam rejecting floating solar still for low-cost
27. Gan, Q., Zhang, T., Chen, R., Wang, X., and Ye,
Generation by Heat Localization. Adv. Mater. desalination. Energy Environ. Sci. 11, 1510–
M. (2019). Simple, Low-dose, durable, and
27, 4302–4307. 1519.
carbon-nanotube-based floating solar still for
efficient desalination and purification. ACS 39. Li, C., Jiang, D., Huo, B., Ding, M., Huang, C.,
16. Li, T., Liu, H., Zhao, X., Chen, G., Dai, J., Pastel, Sustain. Chem. Eng. 7, 3925–3932.
G., Jia, C., Chen, C., Hitz, E., Siddhartha, D., Jia, D., Li, H., Liu, C., and Liu, J. (2019). Scalable
et al. (2018). Scalable and highly efficient and robust bilayer polymer foams for highly
28. Liu, Z., Song, H., Ji, D., Li, C., Cheney, A., Liu, Y., efficient and stable solar desalination. Nano
mesoporous wood-based solar steam Zhang, N., Zeng, X., Chen, B., Gao, J., et al.
generation device: localized heat, rapid water Energy 60, 841–849.
(2017). Extremely cost-effective and efficient
transport. Adv. Funct. Mater. 28, 1707134. solar vapor generation under nonconcentrated 40. Gong, B., Yang, H., Wu, S., Xiong, G., Yan, J.,
illumination using thermally isolated black Cen, K., Bo, Z., and Ostrikov, K. (2019).
17. Hu, X., Xu, W., Zhou, L., Tan, Y., Wang, Y., Zhu,
paper. Glob. Chall. 1, 1600003. Graphene array-based anti-fouling solar
S., and Zhu, J. (2017). Tailoring graphene
oxide-based aerogels for efficient solar steam vapour gap membrane distillation with high
29. Peng, G., Ding, H., Sharshir, S.W., Li, X., Liu, H., energy efficiency. Nano-Micro Lett. 11, 51.
generation under one sun. Adv. Mater. 29,
Ma, D., Wu, L., Zang, J., Liu, H., Yu, W., et al.
1604031.
(2018). Low-cost high-efficiency solar steam 41. Kuang, Y., Chen, C., He, S., Hitz, E.M., Wang, Y.,
18. Wang, G., Fu, Y., Guo, A., Mei, T., Wang, J., Li, generator by combining thin film evaporation Gan, W., Mi, R., and Hu, L. (2019). A high-
J., and Wang, X. (2017). Reduced graphene and heat localization: Both experimental and performance self-regenerating solar
oxide–polyurethane nanocomposite foam as a theoretical study. Appl. Therm. Eng. 143, 1079– evaporator for continuous water desalination.
reusable photoreceiver for efficient solar steam 1084. Adv. Mater. 31, e1900498.
generation. Chem. Mater. 29, 5629–5635.
30. Shi, L., Wang, Y., Zhang, L., and Wang, P. 42. Xu, N., Li, J., Wang, Y., Fang, C., Li, X., Wang,
19. Li, Y., Gao, T., Yang, Z., Chen, C., Luo, W., (2017). Rational design of a bi-layered reduced Y., Zhou, L., Zhu, B., Wu, Z., Zhu, S., and Zhu, J.
Song, J., Hitz, E., Jia, C., Zhou, Y., Liu, B., et al. graphene oxide film on polystyrene foam for (2019). A water lily-inspired hierarchical design
(2017). 3D-printed, all-in-one evaporator for solar-driven interfacial water evaporation. for stable and efficient solar evaporation of
high-efficiency solar steam generation under 1 J. Mater. Chem. A Mater. Energy Sustain. 5, high-salinity brine. Sci. Adv. 5, eaaw7013.
sun illumination. Adv. Mater. 29, 1700981. 16212–16219.
43. Li, X., Ni, G., Cooper, T., Xu, N., Li, J., Zhou, L.,
20. Xue, G., Liu, K., Chen, Q., Yang, P., Li, J., Ding, 31. Fang, J., Liu, Q., Zhang, W., Gu, J., Su, Y., Su, Hu, X., Zhu, B., and Zhu, J. (2019). Measuring
T., Duan, J., Qi, B., and Zhou, J. (2017). Robust H., Guo, C., and Zhang, D. (2017). Ag/diatomite conversion efficiency of solar vapor generation.
and low-cost flame-treated wood for high- for highly efficient solar vapor generation Joule 3, 1798–1803.

14 Cell Reports Physical Science 2, 100330, February 24, 2021

You might also like