You are on page 1of 36

Review

Challenges and Opportunities


for Solar Evaporation
Chaoji Chen,1,2 Yudi Kuang,1,2 and Liangbing Hu1,*

Solar evaporation is an ancient technology that has regained tremendous attention Context & Scale
because of the abundance of solar energy, widely available water sources, and facile Solar evaporation is an attractive
facilities in combination with substantial improvements of conversion efficiency technology that combines the two
enabled by improved photothermal materials, thermal management, and interfacial most abundant resources on
heating system designs in recent years. In this review, we discuss recent develop- Earth: solar energy and water. It
ments in photothermal materials, with a focus on their photothermal conversion has enabled an array of emerging
mechanisms as light absorbers. We also explore the diverse structural design and en- applications, including
gineering strategies that are being used to improve evaporation performance, contaminated water purification,
including the design principles for high-efficiency light-to-heat conversion, optimiza- sea water desalination, electric
tion of thermal management, water transport, interface wettability, and anti-salt- generation, steam sterilization,
blocking structures. We describe the potential applications of this attractive technol- and fuel production. Nonetheless,
ogy in a variety of energy and environmental fields. The current challenges and traditional solar evaporation
future research opportunities are also discussed, providing a roadmap for the future approaches generally heat the
development of solar evaporation technology. entire amount of water in the
system reservoir, leading to a low
Introduction thermal efficiency of just 40%.
Solar energy and water are two of the most abundant resources on Earth.1–4 Solar The recent development of
evaporation, which combines these resources, is regarded as one of the most attrac- interfacial solar evaporation has
tive and facile implementations of solar thermal technology and has been used to enabled just the air-liquid
generate potable water since ancient times.4–7 Steam and clean water generation, interface to be heated rather than
from either wastewater or seawater, is the basic application of the solar evaporation the bulk water, resulting in a much
technique, which represents one of the most promising green and sustainable solu- higher thermal efficiency of up to
tions to the pressing global challenge of water shortages.8–12 Nowadays, more and 90% at reduced solar
more applications are being driven by this fundamental photothermal process, concentration, mainly enabled by
including electricity generation,13 steam sterilization,14 and fuel production.15 How- the rapid development of new
ever, traditional solar evaporation approaches generally deliver a low photothermal photothermal materials and
conversion efficiency of 30%–45% because of poor solar absorption and large heat photothermal structural
losses that result from the placement of the light absorber at the bottom of the water engineering.
source, which in turn greatly hinders the practical application of this technology.16,17
In this review, we highlight recent
Great efforts have been made to improve the photothermal efficiency by enhancing advances in solar evaporation
light absorption and minimizing heat loss.18–26 Halas et al.18 and Deng et al.19 devel- materials, structures, and systems.
oped a volumetric heating system by dispersing Au nanoparticles or light-scattering We discuss different
polystyrene nanoparticles in the bulk water source. In this type of design, light ab- photothermal materials along
sorption can be greatly improved; however, the heat loss is still significant because with their photothermal
the entire system reservoir gets heated in the process. The recent development of mechanisms, as well as various
interfacial solar evaporation systems that place the light absorber at the water-air substrates featuring different
interface has enabled just the air-liquid interface to be heated rather than bulk water, water transport and/or thermal
resulting in a substantially improved photothermal efficiency.27–34 insulation functions, aiming to
provide guidance for future
A typical double-layered interfacial solar evaporation system contains several com- material choices. Diverse material
ponents, including light absorber, substrate, bulk water reservoir, incident light, and

Joule 3, 683–718, March 20, 2019 ª 2018 Published by Elsevier Inc. 683
and structural engineering
strategies toward better solar
evaporation performance are
systematically reviewed, as well as
their emerging applications in
various fields of water and energy
conversion. Finally, we propose
insights into current challenges
and future perspectives of solar
evaporation.

Figure 1. The Overall Material/Structural Scheme of a Typical Double-Layered Interfacial Solar


Evaporation System, Including Light Absorber, Substrate, Bulk Water, Incident Light, and Vapor
The overall design principles and material and structural engineering strategies are also
demonstrated.

vapor (Figure 1). Incident light is absorbed by the light absorber and converted into
heat. In parallel, water is absorbed by the substrate and transported up to the evap-
orative surface through interconnected water pathways via capillary forces. The
generated heat raises the temperature of the water on the evaporative surface,
which drives the evaporation process continuously with the unremitting supply of
water. However, part of the generated heat is inevitably lost to the surrounding envi-
ronment via conduction to bulk water and air, radiation to air, and convection to bulk
water, resulting in an evaporation efficiency lower than 100%. The evaporation effi-
ciency can be determined by the ratio of the stored thermal energy in the generated
vapor to the incoming solar flux, which can be calculated based on the following
relationship11,27:

_ LV Copt P0 ;
h = mh

in which m _ refers to the mass flux (evaporation rate) of water, hLV refers to the total
liquid-vapor phase-change enthalpy (including the sensible heat), P0 is the nominal
solar irradiation value of 1 kW,m 2, and Copt represents the optical concentration.

In the past decade, the efficiency of interfacial solar evaporation has increased up to
90% at reduced solar concentration, mainly enabled by the rapid development of
new photothermal materials and photothermal structural engineering.6,35–41 Mod-
ern nanotechnologies in materials science have driven the development of such
photothermal materials that are able to harvest the full solar spectrum and convert
it into heat with extremely high efficiency. In addition, photothermal structures en-
gineered to minimize heat loss have evolved in parallel. With this progress, interfa-
cial solar evaporation has the potential to drive diverse energy-related applications
in modern society. In particular, interfacial solar evaporation offers clean solar en-
ergy and zero greenhouse gas emissions, providing appealing advantages in ad-
dressing clean energy needs, water scarcity, and global warming issues.

In this review, we outline recent developments in this ubiquitous technology, 1Department of Materials Science and
including the material choice of various types of solar absorbers (e.g., carbonaceous Engineering, University of Maryland, College
materials, plasmonic nanoparticles, semiconductors, and polymers) and the sup- Park, MD 20742, USA
2These
porting substrates, structural engineering strategies to enhance thermal efficiency, authors contributed equally
diverse potential applications, and some key challenges and opportunities that *Correspondence: binghu@umd.edu
require further exploration (Figure 2). We first explore different materials used as https://doi.org/10.1016/j.joule.2018.12.023

684 Joule 3, 683–718, March 20, 2019


Figure 2. Schematic View of the Structure of This Review Article
Material considerations, structural engineering, and applications in the field of solar evaporation. Electricity generation design is reproduced from
Zhou et al., 13 with permission. Heat localization design is reproduced from Chen et al.,30 with permission. 2D water pathway design is reproduced from
Zhu et al., 39 with permission. Biomimetic design is reproduced from Hu et al., 40 with permission. Interface engineering design is reproduced from Zhu
et al., 41 with permission.

solar absorbers and substrates, with a focus on the photothermal conversion mech-
anism of different types of solar absorbers. We then go on to discuss diverse strate-
gies of structural engineering, including high-efficiency light absorption and light-
to-heat conversion; thermal management with the capability of localizing heat on
the evaporative surface; one-, two-, and three-dimensional (1D, 2D, and 3D) water
pathway designs; interface engineering; biomimetic structural designs; 3D solar
absorber design; salt-rejection structural design; and dynamic simulations. The
emerging applications of this intriguing technology in seawater desalination and

Joule 3, 683–718, March 20, 2019 685


wastewater purification, electric generation, sterilization, and chemical fuel produc-
tion are also highlighted. Finally, we describe recent research challenges/opportu-
nities and provide concluding remarks on the future directions of solar evaporation
in both academic research and practical applications.

Materials Choice
Photothermal materials (as solar absorbers) and thermal-insulating and water-trans-
porting materials (as substrates) are key elements in solar evaporation devices. The
performance of solar evaporation is mainly determined by the properties of these
materials. For solar absorber design, high light absorption across the full solar spec-
trum range of 0.3 to 2.5 mm and high light-to-heat conversion efficiency are equally
important, while for substrate material design, good thermal insulating and water
transport properties are desired. The past decade has witnessed a fast growth of
photothermal and substrate materials, mainly enabled by advances in nanotech-
nology, through either new material development or structural engineering of exist-
ing materials. In this section, we will outline recent progress in developing solar
absorber and substrate materials and discuss the photothermal conversion mecha-
nisms of various types of photothermal materials and the design principles of high-
performance solar absorbers and substrates.

Solar Absorbers
Photothermal materials with high light absorption over a wide range of the solar
spectrum have been intensively investigated in the past decade.42–45 Photother-
mal materials are able to absorb incident light and convert it into heat partially
or completely via photoexcitation, in which mobile carriers are driven by the
light-induced electric field and gain energy, which turns into thermal energy
(heat).46 This photothermal effect can be widely observed in various materials,
such as plasmonic metals,47–49 semiconductors,50–53 carbonaceous materials,54,55
and polymers56,57 and their hybrids.58 There are three categories of photothermal
mechanisms; namely, plasmonic localized heating, electron-hole generation and
relaxation, and thermal vibration of molecules. Generally, an individual photother-
mal material converts light to heat via a single mechanism. However, in some
cases, more than one photothermal mechanism may be involved. This is particu-
larly true for hybrid photothermal compounds comprising two or more compo-
nents that demonstrate different photothermal mechanisms. Here we focus on
the discussion of general cases, including plasmonic particles featuring a plas-
monic localized heating mechanism, semiconductors demonstrating an electron-
hole generation and relaxation mechanism, and carbonaceous or polymeric mate-
rials with a photothermal mechanism based on the thermal vibration of molecules.
Thereafter, we offer a generalized discussion and comparison between the
different materials.

Plasmonic Particles
Some metal nanoparticles demonstrate strong light absorption and light-to-heat
conversion capabilities that result from plasmon resonance. Plasmon resonance oc-
curs when the photo frequency matches the natural frequency of electrons on the
surface of the metal. The match between the frequency of the incident light and
the oscillation frequency of delocalized electrons in the metal triggers a collective
excitation of the electrons and the subsequent generation of hot electrons. The
excited hot electrons oscillate coherently with the incident electromagnetic field, re-
sulting in the generation of heat via a joule mechanism (Figure 3A).35,47,59 The hot
carriers redistribute their energy quickly via the electron-electron scattering process,
which heats the plasmonic element itself.

686 Joule 3, 683–718, March 20, 2019


Figure 3. Photothermal Mechanisms of Various
Types of Photothermal Materials (Light
Absorbers)
(A–C) Plasmonic heating (A), electron-hole
generation and relaxation (B), and thermal
vibration of molecules (C).

Many studies have been dedicated to exploring solar evaporation performance us-
ing plasmonic nanoparticles. For example, Halas et al. pioneered the use of gold
(Au) nanoparticles as an efficient solar absorber for solar evaporation,18,21,22 which
inspired many further studies in this field.28,60–67 A self-standing film made by
depositing Au nanoparticles on rough paper demonstrated light absorption of
85% at a wavelength range of 400–800 nm because of the combination of the plas-
mon resonance effect and the multi-scattering effect imparted by the rough cellu-
lose structure.67 In a more recent study by Zhu and co-workers,60 Au nanoparticles
with random size distribution were directly deposited onto a multi-channeled anodic
aluminum oxide (AAO) structure to achieve an ultra-high light absorption of 99%
over a wide wavelength range of 0.4–10 mm. Other metallic nanoparticles, such as
aluminum (Al), silver (Ag), germanium (Ge), and Au-Ag bimetals have also been
investigated as efficient light absorbers for solar evaporation.11,66,68,69

Semiconductors
Recently, low-cost and low-toxicity semiconductors have emerged as a new kind of
photothermal material for solar evaporation. In metal oxide-based semiconductors,
electron-hole pairs are generated when light with an energy similar to that of the
bandgap is absorbed by the material.7,53,70 In narrow bandgap semiconductors,
as the energy of most photons from the incident solar light is higher than that of
the bandgap, above-bandgap electron-hole pairs will be generated.53 These
above-bandgap electron-hole pairs will then relax to the band edges, converting
the extra energy into heat through a thermalization process (Figure 3B). In sharp
contrast, most of the absorbed light energy re-emits as photons in broad bandgap
semiconductors when the electron-hole pair recombines near the bandgap edge,
resulting in a much lower light-to-thermal conversion efficiency.

Narrow bandgap semiconductors, such as hydrogenated black titania,71 Ti2O3


nanoparticles,53 oxygen-deficient MoO3 quantum dots,72 and Fe3O4,73 have also
been investigated as solar absorbers for solar evaporation in the past few years.
Bimetal oxides, such as MnFe2O4, CoFe2O4, and ZnFe2O4 have also demonstrated

Joule 3, 683–718, March 20, 2019 687


excellent photothermal effects when made into free-standing and self-floating
films.73 Most recently, Wang and co-workers74 studied the photothermal properties
and solar evaporation performance of a variety of bi- and tri-metal oxides with a 3D
light absorber design, in which both high light absorption and thermal efficiency can
be achieved.

Carbonaceous and Polymeric Materials


Carbon materials and some polymers demonstrate strong light absorption capabil-
ities (causing them to appear black), being capable of converting this energy into
heat through lattice vibrations. In such materials, abundant, loosely held electrons
can be easily excited from the p orbital to p* orbital with a small energy input.
The excited electron will be promoted from the ground state (highest occupied mo-
lecular orbital, HOMO) to a higher energy orbital (lowest unoccupied molecular
orbital, LUMO) when the photon energy of the incident light matches a possible
electronic transition within the molecule, causing heat to be released when the
excited electron relaxes back to its ground state (Figure 3C).7,75

Because of their high light absorption across a wide range of wavelengths, relatively
low cost, and high stability, various carbon materials have been investigated as light
absorbers for solar evaporation, such as biomass-derived amorphous carbon,76 gra-
phene,77–84 graphene oxide (GO)/reduced graphene oxide (rGO),39,85–90 carbon
nanotubes (CNTs),91–93 graphite,30,94–96 carbon black,97 etc. In addition, carbon ma-
terials can be easily fabricated into various structures desirable for enhanced light
absorption and integrated with diverse substrate materials and evaporation struc-
tures. All these properties position carbon as the most promising photothermal ma-
terial candidate for practical applications. Polymers, although more limited in
choice, also have utility in solar evaporation applications. For example, polypyrrole
(PPy),56,98 polydopamine,99 poly(1,3,5-hexahydro-1,3,5-triazines),100 and polyvinyl
alcohol (PVA)/PPy hybrid hierarchically nanostructured gel57 have recently been
applied as solar absorbers for high-efficiency solar evaporation.

Among these photothermal materials, carbonaceous material is the most competi-


tive candidate in terms of cost, abundance, scalability, stability, and environmental
friendliness. For example, the most abundant and renewable biomass (e.g., trees,
mushrooms, grass, cotton, and other plants) can be easily converted into hard car-
bon via a simple carbonization process. Carbonaceous material is considered the
most practical choice and widely explored for solar evaporation in diverse applica-
tions, including desalination, power generation, and fuel production. Other photo-
thermal materials, such as plasmonic metallic nanoparticles and semiconductors,
have their own sets of bottlenecks toward practical application. For example, the
cost of plasmonic metallic nanoparticles is usually too high, and the fabrication scale
is limited. In addition, the instability and toxicity of these materials are potential con-
cerns for green applications. Polymers, especially polymer hydrogels, demonstrate
some competitive features, such as bonding with water to facilitate the evaporation
process. However, challenges in terms of stability and scalability still remain that
require further exploration.

Substrates
In addition to the light absorber, the substrate material is equally important in high-
efficiency solar evaporation design. The substrate material generally needs to pro-
vide two basic, key functions: water transport/evaporation and thermal insulation.
For efficient water transport and evaporation, both good wetting properties and
continuous pathways are critical. For improved thermal insulation, low thermal

688 Joule 3, 683–718, March 20, 2019


conductivity and porous structures are desirable. It is worth noting that pore struc-
ture is a key element for both water transport and thermal insulation; however, the
balance between efficient water transport (which generally requires pores to be
filled with water) and thermal insulation (which requires pores to be filled with air)
must be considered. Researchers have developed designs to address this conflict
by separating the water-transport pathways from the thermal insulation structures
via 1D or 2D water pathway designs, which will be discussed in more detail in the
next section.

Based on these guidelines for improved water transport and thermal management,
many hydrophilic, porous, and/or thermal insulating materials have been studied as
substrates for solar evaporation. Cellulose foam,31,101 for example, demonstrates
excellent hydrophilicity, low thermal conductivity, and abundant hierarchical
porosity, ensuring sufficient water transport/evaporation and good thermal man-
agement capabilities for efficient solar evaporation. Natural wood, with the majority
of its components being cellulose, possesses good hydrophilicity, well-aligned
hierarchical channels, and low thermal conductivity and, therefore, has inspired
several solar evaporation structures by using wood itself as a substrate mate-
rial.40,66,76,88,91,94,102–104 Polyurethane (PU) and polystyrene (PS) foams, as excellent
commercial thermal insulators, have also been widely used as substrates in solar
evaporation and are also beneficial as they enable interfacial solar evaporators to
easily float on the surface of water.39,78,87,89,105 Other efficient substrate materials
include AAO membranes,11 GO aerogel,86 air-laid paper,67 and carbon nanotube
arrays.106 In general, the rapid development of substrate materials with desirable
water transport/evaporation and thermal insulation properties and diverse structural
designs have boosted the development of solar evaporation devices and systems.

Material and Structural Engineering toward High Evaporation Performance


The past decade has witnessed a substantial improvement in solar evaporation per-
formance, with evaporation efficiency increasing from 30% to 45% to 90% under 1
sun irradiation. Material and structural engineering has played a critical role in this
progress. Essentially, two key design principles are needed for superior solar evap-
oration performance: (1) achieving high light absorption across a wide range of
wavelengths via either material or structural engineering and (2) realizing high-effi-
ciency light-to-heat conversion via material engineering. As described earlier, to
achieve a high evaporation efficiency, good thermal insulation and efficient water
transport are equally important factors in substrate design. Particularly, for the ther-
mal structure design, the key principles are to minimize the heat loss and localize the
heat on the evaporative surface. In this section, we will discuss material and structural
engineering strategies for high evaporation performance, as well as outline recent
progress, aiming to gain insights into the structure-property-performance relation-
ship to guide future material and structural design.

Light Absorption and Light-to-Heat Conversion


Material and structural design enabling high light absorption over the whole solar
spectrum and high-efficiency light-to-heat conversion is the first step toward a
high-performance solar evaporation device. As shown in Figure 4A, the distributed
solar energy on the Earth’s surface covers a wide wavelength range of 300 nm to
2.5 mm, which comprises three parts: the ultraviolet region (300–400 nm, accounting
for 3% of the total energy), the visible light region (400–700 nm, accounting for
45% of the total energy), and the infrared region (700 nm–2.5 mm, accounting
for 52% of the total energy). An ideal light absorber material and structural design
should enable maximum absorption of solar energy across the entire solar spectrum.

Joule 3, 683–718, March 20, 2019 689


Figure 4. Representative Strategies for Tuning the Light Absorption and/or Light-to-Heat Conversion of Photothermal Materials
(A) Solar spectral irradiation across a wavelength range of 300 nm to 2.5 mm, which comprises ultraviolet light (300–400 nm), visible light (400–700 nm),
and infrared light (700 nm–2.5 mm).
(B–D) Tuning light absorption behavior by (B) microstructure tuning, (C) band structure tuning, and (D) particle size and shape tuning.
(C) was modified from Gao et al., 7 with permission.

Various material and structural engineering strategies have been proposed to effec-
tively enhance the light absorption and/or light-to-heat conversion in versatile pho-
tothermal materials, which has been comprehensively summarized in recent review
articles.7 Aiming to provide a concise overview of the state-of-the-art of this exciting
field, here, we mainly focus on some representative strategies, including (1) tuning
the microstructure of the light absorber to enable multi-scattering of incident light
and minimize transmittance and reflection; (2) tuning the band structure of the semi-
conductors via atomic doping to enable light absorption across a wider wavelength
range and higher-efficiency light-to-heat conversion; and (3) tuning the size and/or
shape of plasmonic nanoparticles over a wide range of distribution to enable multi-
ple wavelength light absorption (Figures 4B–4D).

The first strategy is the most popular and universally available for all kinds of photo-
thermal materials. Thus, it can be used to tune the light absorption of photothermal
materials, alone or in combination with the other two strategies. Several recent
studies have demonstrated that structural engineering, either with rough surfaces
or straight pores (i.e., channels), can enhance light absorption through multi-scat-
tering effects.40,67 As discussed previously, the light absorption and light-to-heat
conversion efficiency of a semiconductor are strongly correlated to its band struc-
ture. Tuning the band structure with a narrower bandgap energy by element doping
is beneficial for suppressing reemission of the absorbed light energy through the
recombination of electron-hole pairs near the bandgap edge. In this case, the ab-
sorption wavelength can be broadened to enhance the light absorption of the semi-
conductor (Figure 4D). Chen et al.53 demonstrated an elegant example by tuning the
bandgap of titanium oxide to as narrow as 0.09 eV to enable a higher light absorp-
tion of 92% over the entire solar spectrum.

690 Joule 3, 683–718, March 20, 2019


Figure 5. Thermal Scheme Including Input Power from Incident Light and Output Powers via
Reflection, Thermal Conduction, Radiation, and Convection

For plasmonic nanoparticles, the light absorption behavior, especially the absorp-
tion wavelength, highly depends on the size and shape of the particle. Narrow
size-distributed and single geometric plasmonic nanoparticles only demonstrate a
limited absorption range within a few wavelengths. By tuning the size and/or shape
of plasmonic particles, light absorption across a broader range can be achieved (Fig-
ure 4D). In a recent study, Zhu and co-workers demonstrated that pronounced
broadband absorption can be obtained by using closed-packed random gold
nanoparticle assemblies in a large pore-sized absorber, which is dominated by
high density of the optical modes (provided by random gold nanoparticles) and
strong multiple scattering (caused by the nanoporous template).60

Heat Localization and Thermal Concentration


Thermal management is critical in solar evaporation systems. The energy balance
scheme demonstrated in Figure 5 indicates how material and structural designs
can provide high-efficiency thermal management and vapor evaporation by mini-
mizing thermal loss caused by conduction, convection, and radiation. There are
two general design principles for thermal structure design: (1) minimizing heat
loss via material (including solar absorber and substrate material) and thermal struc-
ture in both microscopic scale and macroscopic scale; (2) localizing the generated
heat on the evaporative surface to enable a high surface temperature and fast evap-
oration rate. In the past decade, tremendous efforts have been dedicated to
improving the thermal management capability of solar evaporation devices, which
in turn substantially enhances the evaporation efficiency.

The heat localization design by Chen and co-workers represents a brilliant thermal
structural design30 in which they utilized a double-layer carbon structure combining
absorbing, hydrophilic, porous exfoliated graphite as the light absorber, and insu-
lating, hydrophilic, porous carbon foam as the thermal insulator (Figures 6A–6C).
The two layers of the carbon structure demonstrate different thermal properties (Fig-
ures 6D–6G). The top layer of exfoliated graphite has a relatively high thermal con-
ductivity of 0.309 W m 1 k 1, which will further increase to 0.959 W m 1 k 1 if wetted
by water. In contrast, the bottom layer of the carbon foam exhibits an extremely low
thermal conductivity of 0.117 W m 1 k 1 at a dry state, mainly due to its highly
porous structure with both open and closed pores. When wetted by water, the ther-
mal conductivity of the carbon foam significantly increases to 0.426 W m 1 k 1,

Joule 3, 683–718, March 20, 2019 691


Figure 6. High-Performance Solar Evaporation via Heat Localization
(A–C) Double-layer structure of a solar evaporation device. (A) A representative structure for localization of heat. The cross-section demonstrates the
temperature distribution in the device. (B) A double-layer structure that consists of carbon foam supporting an exfoliated graphite layer. Both layers are
hydrophilic to promote the capillary rise of water to the surface. (C) A picture of enhanced steam generation by the double-layer structure under a solar
illumination of 10 kW m 2 .
(D–G) Thermal conductivity of the double-layer structure: (D) exfoliated graphite in air, (E) carbon foam in air, (F) exfoliated graphite with water, and (G)
carbon foam filled with water. The insets are representative images taken by an IR microscope.
(H) While the steam is generated, the underlying bulk liquid is at the ambient temperature.
(I and J) The evaporation mass loss of water with different structures under 1 and 10 kW m 2 solar irradiation: (I) 1 sun, (J) 10 sun.
(K) The solar thermal efficiency and the corresponding evaporation rate at various optical concentrations. Reproduced from Chen et al., 30 with
permission.

mainly because of the transformation of open pores from air filled to water filled.
Because of the better thermal insulating property, the majority of generated heat
can be blocked by the carbon foam layer from dissipating into the bulk water by con-
duction. As shown in Figure 5B, this design resulted in the enhanced temperature of
the generated vapor phase and underlying cold water under 10 kW m 2 due to the
localized energy close to the surface of the structure accompanied by a limited flow
rate of the fluid. In this case, heat can be mainly localized on the evaporative surface,
contributing to a substantially higher evaporation rate than bulk water and volu-
metric absorber with only exfoliated graphite (Figures 6I and 6J). Consequently, a
high solar thermal efficiency of up to 85% at only 10 kW m 2 can be achieved
(Figure 6K).

This heat localization structural design has inspired many further studies.107–113
Some recent studies have further developed this concept by totally separating the
thermal insulation structure from the water transport structure, which prevents water
from filling the open pores of the thermal insulation layer, resulting in a better heat
localization effect and higher evaporation efficiency.39,114–116

692 Joule 3, 683–718, March 20, 2019


Thermal concentration is another smart strategy for the direct utilization of ambient
solar flux to generate vapor, in which the generated heat from a large area of light
absorber can be concentrated within a small evaporative area.117 In a recent study,
researchers combined such a thermal concentration structure with a thermal insula-
tion structure to achieve an exceptionally high vapor temperature of up to 100 C,
even under ambient solar flux.117 It is worth noting that although optical concentra-
tors have a similar effect by increasing the incoming light density on the evaporative
surface, the need for additional facilities would add greater complexity and cost to
the system, which may hinder its wide application, especially in remote areas. Ther-
mal concentrator designs that do not require additional facilities would provide a
better choice for utilizing natural solar flux in a more cost-effective and easier
manner.

3D, 2D, and 1D Water Pathway Designs


Water pathway design is a key factor for achieving continuous and efficient water
transport. In traditional solar evaporation devices, 3D random and interconnected
porous structures enable the continuous transport of water from the underlying
reservoir to the upper evaporative surface. In a recent study,103 a nature-made
3D interconnected porous structure was demonstrated via a reverse-tree design
by placing the wood channels parallel to the water surface (Figures 7A and 7B).
These 3D water-pathway designs with abundant and open pores facilitate water
transport with sufficient capillary effect as a driving force. However, the filling of
water in open pores always leads to increased thermal conductivity, which causes
more heat loss through conduction to the bulk water and weakens the heat local-
ization effect.

Recently, Zhu’s group39 developed a 2D water pathway structure by confining water


transport in a 2D cylindrical surface wrapped on a PS foam thermal insulator (thermal
conductivity: 0.04 W m 1 K 1), enabling both efficient water supply and sup-
pressed heat loss simultaneously. The GO film absorber in this design is physically
separated by a thermal insulator instead of being in direct contact with bulk water
to greatly suppress parasitic heat loss (Figures 7C and 7D). A thin layer of hydrophilic
cellulose was wrapped over the surface of the thermal insulator, providing a 2D path
for water transport pumped by capillary force (Figures 7E and 7F). Compared with
‘‘bulk water supply’’ in conventional designs, the confined 2D water pathway struc-
ture reduces the dimensionality and volume of the water path, thus minimizing the
heat dissipation to the bulk water through conduction (Figures 7C and 7D). In a
more recent study, the same group further reduced the dimensionality of the water
path structure from 2D to 1D to further suppress heat loss (Figures 7G and 7H).115 An
angular absorber combined with the 1D water path design successfully minimized
radiation, convection, and conduction losses, enabling over 85% solar evaporation
efficiency under ambient solar flux, without external thermal insulation and optical
supporting systems. In the same year, Li et al.87 combined a 1D water path structure
with a thermal insulating material (expanded polystyrene) in a jellyfish-like solar
evaporator via a fast and facile 3D printing manufacturing approach. The resulting
evaporator demonstrated an impressive energy conversion efficiency of 87.5% un-
der 1 sun illumination due to the 1D water path and thermal insulating structural
design.

One common feature of 1D and 2D water pathway designs is the separation of the
thermal insulation structure from the water transport structure. The pores in the ther-
mal insulator are designed to be closed in order to minimize heat loss, while the
pores in the water transport structure are open to enable sufficient and continuous

Joule 3, 683–718, March 20, 2019 693


Figure 7. 3D, 2D, and 1D Water Path Designs
(A and B) 3D water path design enabled by a horizontally placed porous wood structure, in which the abundant lumina and pits constitute a 3D
interconnected porous structure for efficient water transport: (A) schematic illustration, (B) SEM images.
(C–F) 2D water path design by confining the water to the outer surface of a cylindrical thermal insulator: (C) schematic of conventional solar evaporation with direct
water contact, (D) schematic of solar desalination devices with suppressed heat loss and 2D water supply, (E) flowchart for the fabrication of solar desalination
devices, featuring polystyrene foam, a cellulose coating, and a GO film on the top surface, and (F) physical map of the polystyrene foam (thermal insulator, left),
cellulose (2D water path, middle) wrapped over the surface of the polystyrene foam, and the GO film (absorber, right) on the top surface.
(G and H) Schematics of conventional and 1D water path designs: (G) conventional, (H) 1D water path.
(A) and (B) were reproduced from Hu et al., 103 with permission; (C)–(F) were reproduced from Zhu et al., 39 with permission; (G) and (H) were reproduced
from Zhu et al., 115 with permission.

transport of water. In this way, the conflicting needs of water transport and thermal
insulation can be balanced.

Interface Engineering
Interfacial properties play another important role in solar evaporation systems. The
most important property of the interface is its wettability, which is important for
water transport, light absorption, device floating, and stability. A hydrophilic inter-
face (including the surface) is desirable for wicking water through capillary pumping,
which has been a common design principle for absorber and water transport struc-
tures. However, some recent studies have suggested that not all hydrophilic inter-
faces are good for high-efficiency water evaporation, especially for the top evapora-
tive surface.56 Being too hydrophilic on the top evaporative surface sometimes leads
to excess water covering the top surface, inducing more heat loss and making it diffi-
cult to maintain self-sustained solar evaporation.

To address this issue, in 2015 Wang et al.56 demonstrated a robust and self-healing hy-
drophobic evaporative structure by using a fluoroalkylsilane modified PPy-coated stain-
less steel (SS) mesh film (Figures 8A–8D). The surface wettability of the film was

694 Joule 3, 683–718, March 20, 2019


Figure 8. Interface Engineering Strategies in Solar Evaporation
(A–E) Hydrophobic light-to-heat conversion membranes. (A) Schematic illustration for the preparation of the light-to-heat conversion membrane for
interfacial heating. (B and C) Scanning electron microscopy (SEM) images of the original smooth stainless steel mesh surface and the PPy-coated mesh
surface with microstructures: (B) original mesh, (C) PPy-coated mesh. (D) Dependence of the thickness of the PPy coating on the electropolymerization
time. (E) Water contact angles of the PPy-coated stainless steel meshes after the hydrophobic modification. Insets in (E) show the shapes of a water
droplet on the membrane at different tilt angles of 0  (contact angle about 140  ), 90  , and 180  , showing typical Wenzel’s wetting behavior. Reproduced
from Wang et al., 56 with permission.
(F–H) Janus absorber membrane for solar desalination. (F) Schematic of solar evaporation and desalination and the structure of the Janus absorber. (G)
Flowchart for the fabrication of the Janus absorber, including the fabrication of the PAN and PMMA membranes by sequential electrospinning, followed
by spray deposition of CB nanoparticles. (H) Images of a large-area and curved Janus absorber, demonstrating the washing stability of the Janus
absorber (washed 16 times) and the CB/PAN membrane (washed once). Reproduced from Zhu et al., 41 with permission.

controlled to conform the Wenzel’s wetting behavior, enabling the membrane to spon-
taneously stay at the water-air interface and allow for sufficient contact with the water
surface in order to precisely heat the interfacial water under solar irradiation (Figure 8E).
This hydrophobic photothermal film demonstrated durable and significantly enhanced
water evaporation benefiting from its self-healing hydrophobic structural design. In a
more recent work,41 a hydrophobic top evaporative surface along with a hydrophilic
bottom surface was combined in a single solar evaporation device (Figure 8F). The Janus
membrane consisted of a hydrophobic carbon black (CB)/poly(methyl methacrylate)
(PMMA) top layer working as a light absorber structure and a hydrophilic polyacryloni-
trile (PAN) bottom layer as a water transport structure (Figures 8G and 8H). Taking
advantage of the Janus structure, the light absorber and water pumping structures
can be separately designed with different wettabilities as desired, enabling stable, du-
rable, and high-efficiency solar evaporation.

Joule 3, 683–718, March 20, 2019 695


Following the same guideline, Yan and co-workers118 developed a 2D flexible bilay-
ered Janus membrane with a hydrophobic-assembled gold-nanorod top layer as a
light absorber and a hydrophilic single-wall carbon-nanotube bottom layer as a
wicking and thermal insulating structure, which demonstrated a high evaporation
efficiency of 94% under 5 kW m 2 solar irradiation and stable water-generation
capability during long-term illumination cycles. It is thought that more interface
engineering and designs via diverse strategies such as surface chemistry modifica-
tion,119 surface structural design, and evaporator structural design will boost the
development of solar evaporators with both high efficiency and good stability in
the near future.

Biomimetic Structural Design


Nature is the greatest master at constructing brilliant structures for natural processes
through billions of years of evolution, inspiring many structural designs in a range of
research fields.120–122 Plants, for example, commonly have a multi-channeled struc-
ture that can transport water and nutrients from the roots to the upper trunk, which
are then used during the photosynthesis process. From a material point of view,
cellulose-dominated components of plants possess good hydrophilicity and low
thermal conductivity, which along with the well-developed multi-channeled struc-
ture are all desirable features for a solar evaporation device. The similarity between
the nature-developed structure of plants and solar evaporation structures has
inspired a variety of biomimetic designs toward a high-performance solar evapora-
tion device.

Hu et al. and Singamaneni et al. have contributed several pioneering tree- and reverse-
tree-inspired designs by mimicking the water transpiration process that occurs in trees
using a bilayer wood structure.40,66,76,88,91,94,103,104 In a typical tree-inspired design,40 a
natural basswood membrane is surface-carbonized on a hot plate in air to construct a
thin layer of amorphous carbon on the top surface (Figures 9A, 9B, and 9D). When
placed in seawater or wet sand and illuminated by incident light, the black top surface
acts as light absorber, which quickly heats up and generates vapor (Figures 9C and 9E).
The unique hierarchically aligned porous structure of natural wood123–125 with its good
hydrophilicity ensures the continuous supply of water from the bottom to the upper
evaporative surface via capillary pumping (Figure 9F).126 Meanwhile, the uncarbonized
wood substrate demonstrates a low thermal conductivity (0.2 W m 1 K 1), which can
potentially block the majority of the generated heat from dissipating into the bulk water
and environment. The black and rough carbonized layer with a multi-channeled struc-
ture demonstrates a high light absorption of 99% across the entire solar spectrum.
The systemic advantages of sufficient water transport, good thermal insulation, and
ultra-high light absorption contribute to a high evaporation rate of over 11 kg m 2
h 1, both in seawater and ground water (Figures 9G and 9H). Another intriguing advan-
tage of the bilayer wood-based solar evaporation device is the fact that wood is an
abundant resource with low cost ($1 m 2, much lower than any other existing technol-
ogies), which offers a practical solar energy conversion and water extraction solution
(from seawater, sand, and soil) capable of providing large-scale clean water to remote,
arid, or disaster-relief areas.

In addition to the direct surface carbonization by hot plate in air, other strategies,
such as flammable surface carbonization,102 GO coating,88 CNT coating,91 graphite
coating,94 and plasmonic nanoparticle coating66 have been demonstrated on
bilayer wood-based solar evaporation devices. Wood species-dependent water
transport and thermal insulation properties and behaviors have also been systemat-
ically investigated in a recent work,76 which provides guidelines for the structural

696 Joule 3, 683–718, March 20, 2019


Figure 9. Biomimetic Design of a Bilayer Wood-Based Solar Evaporation Device
(A–C) Schematic illustrations of the tree-inspired design: (A) the water transpiration process, (B) surface carbonized bilayer wood structure pumps water
and generates vapor when illuminated, and (C) the bilayer wood-based device can work in both seawater and ground water.
(D) Digital images of the natural and surface-carbonized wood blocks.
(E) Digital image of a working device.
(F) The hierarchically aligned and porous structure of wood.
(G and H) Evaporation rates of the bilayer wood-based device under various light intensities in seawater and ground water systems: (G) groundwater, (H)
seawater. The performance of bulk water under the same conditions was also included for comparison. Reproduced from Hu et al., 40 with permission.

engineering and optimization of wood-based and other types of solar evaporation


devices and systems. In addition to trees, mushrooms,127 grass,128 and plant
leaves109 have also inspired smart biomimetic structural designs for solar
evaporation and have demonstrated some of the best performances. By analyzing
nature-made structures and learning naturally occurred behaviors more intensively,
researchers can further boost the development of solar evaporation materials and
devices via nature-inspired designs.

Joule 3, 683–718, March 20, 2019 697


Salt-Rejection Structural Design
Even though great improvements have been made in the development of new ma-
terial/structural designs with increasing efficiency, interfacial solar evaporation is still
not ready for practical application, particularly because of its lack of long-term sta-
bility. The most common issue accompanying continuous vapor generation is the
accumulation of salt at the heating interfaces, which creates two serious problems:
(1) decreased solar-thermal conversion caused by the increased solar reflection of
the salt deposit layer and (2) blocking the water supply path, both of which lead
to dramatic decreases in evaporation efficiency. Conventional methods in response
to the salt blockage issues include backwashing or physical removal; however, this
would interrupt continuous production, leading to increased operation costs and
decreased productivity. Novel solar absorber materials with anti-fouling properties
and solar desalination systems with salt rejection properties are therefore highly
desired but also challenging to achieve.

Recently, several methods have been reported to address the salt-blocking issues of
solar evaporation systems.41,66,129–131 For example, inspired by the transpiration ef-
fect of trees, the Hu group66 has reported a bilayer wood evaporator featuring the
inherited structure of its wood parent material, which can work continuously in
seawater without salt deposition when the solar irradiation is below 5 sun. This na-
ture-inspired unique channel design takes advantage of the abundant, large, and
straight open micro-channels in wood for water transport, enabling the re-dissolu-
tion of salt for a self-cleaning solar desalination system. Floating film absorbers
with asymmetric wettability have also been reported with good anti-salt blocking
properties.41 However, because of the increased thermal conduction loss, the so-
lar-evaporation efficiency of these film-like absorbers is relatively lower than the
most promising results reported for solar evaporation.

To maintain the heat localization, Chen’s group129 presented a floating solar still
with a salt-rejecting evaporative design, shown in Figures 10A–10D. In this work,
specially arranged expanded polystyrene with low thermal conductivity (0.02 W
m 1 K 1) was placed under the solar absorption layer to minimize the thermal con-
duction loss to the bulk solution, while simultaneously rejecting salt at the wick
structure via advection and diffusion. Stable solar evaporation performance was
achieved in both freshwater and saltwater (3.5 wt % NaCl), as shown in Figures
10E and 10F. Moreover, the solar still design also demonstrated excellent salt rejec-
tion behavior by dissolving and rejecting salt crystals through the wick under solar
irradiation (Figures 10G and 10H). Despite these progresses, more efforts are
needed in this area to address the salt blocking issue while maintaining a high
evaporation efficiency as well as the long-term stability in harsh conditions, such
as highly concentrated brine and solar irradiation. It should be noted that the salt
blocking phenomenon can be a benefit if producing salt rather than clean water
is the purpose of the device.132

Macroscopic 3D Absorber Design


Recently, nearly or even greater than 100% solar steam efficiency has been reported
by changing the structure of the absorber from a conventional planar device to a 3D
structure at the macroscopic scale, which suggests a new way to improve the solar
steam performance beyond current solar-to-steam limits.74,82,133–137 Using a 3D
structure dramatically increases the evaporation surface area, which provides
more evaporative interfaces per unit area. For example, Kwanghyun et al.82 reported
a mesoporous 3D graphene network/wood evaporator with extended evaporation
sites, which demonstrated an evaporation rate of 1.64 kg m 2 h 1 under 1 sun

698 Joule 3, 683–718, March 20, 2019


Figure 10. A Salt-Rejection Structural Design of Solar Evaporation Device
(A) The evaporator’s structural design, featuring black fabric for solar absorption and a composite white fabric wick and polystyrene foam insulation.
(B) The advection flow of salt rejection resulting from the presence of denser, saltier water at the evaporation surface.
(C) Photograph of the evaporator structure.
(D) Schematic of the evaporator in a fabricated polymer-film-based condensation cover operating in an ocean.
(E) Evaporation rate of the evaporator in fresh (dashed) and salt water (solid, 3.5 wt % NaCl).
(F) Performance of the evaporator in freshwater at different solar fluxes below 1 sun (1 kW m 2 ).
(G) The progression of salt rejection from the evaporator under 1 sun illumination. The evaporator is placed in a reservoir of 3.5 wt % NaCl, and enough
solid NaCl is placed on the evaporator to saturate (26 wt %) the structure.
(H) A separate test to visualize saltwater rejected by the evaporator. Excess salt at the evaporation surface forms a denser solution, which sinks into the
water reservoir. Reproduced from Chen et al., 129 with permission.

irradiation, corresponding to an overall efficiency of 91.8%. The authors claimed that


evaporation rate was higher than that of a 2D planar evaporator control because of
the enlarged side surfaces of the evaporator.

Joule 3, 683–718, March 20, 2019 699


Besides a large specific area, 3D absorbers have also been reported to have the
capability to harvest energy from the environment to enhance the solar evapora-
tion performance. Previous work by Gan and co-workers137 demonstrated that
cold vapor can be generated by rational structural design, which enables the solar
absorber to gain energy from the warmer surrounding environment. As a result, a
near unity conversion efficiency of absorbed solar energy can be realized. More
recently, Zhu’s group133 demonstrated that conventional solar absorbers usually
focus on enhancing the solar-thermal effect to improve the evaporation rate, but
the higher temperature of the absorber will cause energy loss to the environment.
By utilizing a 3D absorber with careful structural design (Figures 11A–11C), the
non-irradiated surface temperature of the evaporator can be lower than the tem-
perature of the surrounding environment due to the evaporation cooling effect,
which allows the absorber to collect energy from the environment to enhance
steam generation (Figure 11D). In this case, the total energy input during the solar
steam process is equal to the incident solar energy plus the thermal energy from
the environment, enabling a greater-than-theoretical limit of solar-to-steam effi-
ciency (Figure 11E).

Similar environmental energy harvesting and over-limit solar vapor generation have
also been reported by Wang’s group,74 who designed a 3D cylindrical absorber with
a cup-shaped structure. Compared with a 2D absorber, the 3D cup structure can effi-
ciently decrease the diffuse reflection energy loss (from 17.9% to 2.5%) and thermal
radiation loss because of multi-scattering and absorption in the cup-like absorber,
therefore leading to an efficient utilization of solar energy and higher evaporation
rate (Figures 11F–11M). Moreover, because of its long pathway before the
generated hot vapor leaves the cup-like absorber, part of the thermal energy in
the hot vapor will be transferred to the cool wall by thermal conduction, convection,
and radiation, which further increases the energy efficiency. These 3D evaporator
designs represent a promising direction for developing high-performance solar
evaporators by minimizing the thermal loss of the system and simultaneously gain-
ing extra energy from the hot vapor and/or surrounding environment.

Theoretical Simulation of Light Absorption, Water Transport, and Thermal Transfer


Theoretical simulation is a powerful technique that can not only help us gain insights
to the properties and transport behaviors of light, water, vapor, and thermal energy
but also guide material and structural design with optimized properties and perfor-
mance for solar evaporation. Note that some behaviors, especially molecular or
atomic-level behaviors, are difficult to characterize by current experimental
methods, making it highly challenging to better understand these properties.
Computational methods, however, can simulate these behaviors using theoretical
models and software tools as an important supplement to experimental analysis.
In the past few years, simulations of light transport and absorption, water
transport/pumping, vapor generation, and thermal management have been
performed to understand how these behaviors depend on different materials and
structures.40,67,93,117,138–140

For example, the influence of microstructure in a carbonized wood absorber was


evaluated via optical absorption simulation in a recent study (Figure 12A).40 The
modeling results indicate that both big and small channels can help improve the
light absorption by multi-scattering. The size of the channels also matters; wood
channels with a larger diameter (up to 50–100 mm) can guide light deeper into the
wood block, while the surface of the smaller channels (diameter z10 mm) predom-
inantly absorb the incident light.

700 Joule 3, 683–718, March 20, 2019


Figure 11. Macroscopic 3D Absorber Design toward Enhanced Solar Evaporation
(A–E) Environmental energy-enhanced interfacial solar evaporation. (A) Comparison between conventional interfacial solar evaporation and
environmental energy-enhanced interfacial solar evaporation. (B) Photograph of the environmental energy-enhanced interfacial solar vapor generator.
(C) Infrared images of the environmental energy-enhanced interfacial solar vapor generator under 120 mW/cm2 . (D) Temperatures of the top and side
surfaces of the environmental energy-enhanced interfacial solar vapor generator under different light intensities. (E) Evaporation rate and enhancement
factor of the environmental energy-enhanced interfacial solar vapor generator under different light intensities. Reproduced from Zhu et al., 133 with
permission.
(F–M) A 3D photothermal structure toward improved energy efficiency. (F and G) Schemes of energy loss pathways of diffuse reflection and thermal
radiation, for photothermal materials featuring 2D disk and 3D cup structures: (F) thermal radiation, (G) diffuse reflection. (H–K) Images of PQC-19
materials composed in (H) 2D flat disk and (I) 3D cup structures in the wet state, and the corresponding IR images (J and K) in the wet state under 1 sun
illumination. (L) Reflectance spectra of the PQC-19 material featuring 2D disk and 3D cup structures in dry and wet states. (M) Evaporation rate recorded
in darkness (R dark ), under light (R light ), and their difference (R net = R light R dark ) for 3D cup-shaped PQC-19 materials with different wall heights.
Reproduced from Wang et al., 74 with permission.

In another example, water transport behavior in a horizontally placed porous wood


structure was numerically modeled with the aim of quantifying the relative role of the
pits and longitudinal channels of the wood structure in dictating the water transport
(Figure 12B).103 In the modeling, the water lost because of evaporation was instan-
taneously replenished by the water that rises to the top of the wood surface by capil-
larity, corresponding to when the evaporation rate is constant. Modeling of this flow
through the wood structure unit showed that water indeed can transport through the
micro-capillaries of the horizontally placed wood, in which the tiny pits present in the
wood structure play an important role in water transport by regulating the flow along
the perpendicular-to-tree growth direction.

Joule 3, 683–718, March 20, 2019 701


Figure 12. Simulations of Light Transport, Water Transport, Vapor Generation, and Thermal
Diffusion Behaviors
(A) Modeling of light transport behavior in a bilayer wood structure. Reproduced from Hu et al., 40
with permission.
(B) Simulations of water transport behavior in a horizontally placed wood structure. Reproduced
from Hu et al., 103 with permission.
(C and D) Simulation of vapor generation in a multilayer graphene membrane. (C) Schematic
diagram for the simulation setup. Multilayer graphene membranes with cone-shaped nanopores
are shown in gray; nitrogen, oxygen, and hydrogen molecules are in blue, red and white,
respectively. (D) The potential, force (along the Z direction), and profile of multilayer graphene with
different A min (from 22.1 Å 2 to 55.2 Å2 ). Reproduced from Yang et al., 138 with permission.
(E and F) Simulation of heat transfer behavior in a thermal concentration structure. (E) Heat transfer
diagram of the perfect conduction model. The selective absorber and evaporator are assumed to
be isothermal. (F) Efficiency of the receiver versus thermal concentration. The dots are
measurements, and the lines are computed using the 1 sun ambient steam generator heat transfer
model. Reproduced from Chen et al., 117 with permission.

Vapor generation is another aspect that requires fundamental understanding, espe-


cially related to the microstructure, thermal gradient, or capillary force. A recent
study138 investigated the fast vapor generation behaviors of a multilayer graphene
membrane with cone-shaped nanochannels via molecular dynamic simulations.
The cone-shaped nanochannels play an important role in driving the fast movement
of water molecules. Molecules can be pumped through these cone-shaped nano-
channels unidirectionally to enhance vapor generation due to the nano-ratchet ef-
fect, which results from the asymmetric force field in the cone-shaped nanopores
(Figure 12C). Several factors, such as the number of graphene layers, cross-sectional
area, cone angle, functional groups, and temperature, may affect the efficiency of

702 Joule 3, 683–718, March 20, 2019


the graphene nano-ratchet. For example, the area of the minimum pore (Amin) will
affect the force distribution of the membrane, that is, the negative (leftward) force
regions expand and the positive (rightward) force regions shrink significantly with
increasing Amin (Figure 12D).

Simulation can also help us understand the thermal behaviors in the solar evapora-
tion system. As demonstrated in a thermal concentration structural design by Chen
and co-workers,117 modeling of the heat transfer behaviors was carried out to gain
insights into the observed experimental results and future performance optimiza-
tions. A simple fin model was used to justify that temperature throughout the selec-
tive absorber is nearly uniform and consistent with the measurements (Figure 12E).
This isothermal assumption was further incorporated into the isothermal model,
along with the performance of COMSOL simulations to determine sidewall losses
in the lab-scale experimental 1 sun ambient steam generator, as plotted in Fig-
ure 12F. The simulated results matched well with experimental observations, which
provides further understanding for future performance. Simulation is a powerful tool
that has played an important role in the understanding of structure-property-perfor-
mance relationships in solar evaporation, helping to guide and optimize material
and structural design, which will continue to play a critical role in future studies.

Multifunction toward Various Applications


The enormous progress that has been made in the field of solar evaporation
provides a great opportunity for boosting the practical application of sustainable
energy and water toward a variety of fields. The mass and energy transfer and trans-
formation between various forms in the solar evaporation system enable this tech-
nology to convert unpurified water (e.g., seawater, wastewater, and contaminated
water) into clean water and solar energy into other forms of energy (e.g., thermal,
chemical, electrical, and mechanical energy) that can be easily collected and stored
for further use. In this section, we will discuss the potential applications of solar
evaporation, including wastewater purification and seawater desalination, electricity
generation, chemical fuel production, and solar steam sterilization.

Desalination/Water Purification
Water and energy scarcity are two major global challenges facing modern society.
Right now, almost one-fifth of the world’s population is living in areas with water
scarcity, and another 1.6 billion people are living in economic water scarcity areas
because of technical or financial limitations to getting freshwater even when water
is available.141 This circumstance will be much more serious by the year 2025,
when it is projected that two-thirds of the world’s population will be under water
stress conditions, according to the United Nations’ World Water Development
Report in 2012 (Figure 13A).142 Thus, a technology that combines the advantages
of freshwater generation, easy accessibility, and cost-efficient energy input are of
great interest for dealing with this global water crisis, especially for people living
in off-grid areas.

Solar energy-enabled water treatment methods are considered attractive for pro-
ducing clean water from a variety of undrinkable water sources, including seawater,
river/lake water, and contaminated water, in a green and sustainable way.143 Partic-
ularly, solar desalination can utilize abundantly available and cost-free solar energy
(Figure 13B) to produce drinkable water from natural water sources, making it a
promising candidate in response to the water scarcity issue,144 as well as being
attractive for its low environmental impact. But conventional solar-still devices

Joule 3, 683–718, March 20, 2019 703


Figure 13. Water and Solar Energy Distribution around the World
(A) Projected global water scarcity in 2025 (IWMI 2000). Adapted from Neto, 142 with permission.
(B) Global horizonal irradiation. Reproduced from Solargis, GIH map,144 with permission.

with low solar-to-vapor efficiency and high capital construction costs significantly
hinder their practical application in off-grid areas.

Recently, with the enormous progress made in photothermal material synthesis and
solar evaporation structure design, great attention has been refocused on the devel-
opment of solar desalination systems with improved solar evaporation efficiency,
miniaturized structures, and affordable costs. For example, Zhu’s group11 reported
a portable floating absorber with a broad solar absorption (96%) for solar desalina-
tion. The absorber was prepared by the self-assembly of Al nanoparticles into a
commercially available 3D porous membrane, as shown in Figures 14A–14F. Upon

704 Joule 3, 683–718, March 20, 2019


Figure 14. Solar Evaporation toward Seawater Desalination
(A–F) Fabrication process and characterization of the Al nanoparticle (NP)-based plasmonic structure: (A) graphical illustration of the fabrication
process, (B) optical photographs of the aluminum foil, anodic aluminum oxide membrane (AAM) sample, and Al NP/AAM structure observed from the
AAM side. (C–F) high-resolution SEM images of the structure: (C) top-view image, (D) cross-sectional view image, (E) magnified top-view image, (F)
magnified cross-sectional view image.
(G) Experimental setup for solar desalination and schematics of the plasmon-enhanced solar desalination process.
(H–K) Desalination performance and the durability of the device. (H) The measured salinities (the weight percentage of Na +) of the four simulated
seawater samples before and after desalination. The dashed colored lines refer to the WHO and EPA standards for drinkable water. (I) The measured
concentrations of four primary ions in an actual seawater sample (from the Bohai Sea, China; average salinity 2.75 wt %) before and after desalination.
The purple (green) shaded area refers to the overall typical salinity achieved by traditional membrane (distillation) desalination processes, respectively,
and (J and K) the evaporation cycle performance of the Al NP/AAM structure under different solar concentrations and under constant illuminations of 2
sun and 4 sun. Each cycle was sustained for 1 h.
Reproduced from Zhu et al., 11 with permission

irradiation, the plasmon hybridization and localized surface plasmon resonance of


the Al nanoparticles lead to localized heating of the absorber, enabling a steady
steam generation of 5.7 kg m 2 h 1 under 4 sun illumination (Figure 14G). The desa-
lination effect of the evaporator was further evaluated by testing in four types of saline
water and real seawater samples with different salt concentrations (0.8–10 wt %). The
results showed that the salinity of the distilled water was approximately four orders of
magnitude lower than the bulk solution, and below the drinkable water standards of
the World Health Organization (WHO) and the US Environmental Protection Agency
(EPA) (Figures 14H). Moreover, the ion concentration in the distilled water was found
to be lower than membrane-based desalination, suggesting the excellent desalina-
tion effect of the absorber (Figure 14I). The durability of the absorber was also
explored under 2 and 4 sun irradiances for 20 cycles, with each cycle illuminated
for 1 h (Figure 14K). Other evaporators in forms of wood membrane, sponge, aerogel,
textile, paper, and multilayer integrated structures have also been reported to show
excellent salt removing effects.40,41,78,82,84,118,129,130,135,145–148

It is worth noting that most of the reported solar desalination experiments were eval-
uated in a steady and open environment, and therefore the evaporation rate cannot

Joule 3, 683–718, March 20, 2019 705


Figure 15. System-Level Integration of Source Water Supply, Solar Evaporation, Vapor
Condensation, and Clean Water Collection

be directly converted to the final water products. In a practical solar desalination sys-
tem, the absorber is usually placed in a sealed space, as shown in Figure 15. In this
case, the total evaporation efficiency is much lower than that in the open environ-
ment because of the increased solar reflection caused by the condensed water
droplet on the cap and the high humidity in the limited space. Meanwhile, the dra-
matic latent heat loss during water vapor condensation also leads to decreasing wa-
ter production. Therefore, more innovations are needed in the system-level design
of solar desalination to further improve the final water yield. For the first challenge,
new material and structural designs are needed to mitigate the reflection of incident
light to enhance the utilization of solar energy in a closed system. For the second
challenge, the combination of interfacial solar heating with membrane distillation
is a potential solution.149,150

Additionally, structure designs enabling multiple cycles of latent heat recovery are
also desirable for mitigating the latent heat waste in the water vapor condensation
process.151–153 In a recent study, Zhou’s group152 developed a solar thermal mem-
brane distillation device combining two layers of distillates with a hydrophobic
membrane to effectively recycle the latent heat with significantly enhanced clean
water productivity and solar efficiency. In another example, Asinari and co-
workers153 demonstrated a passive, modular, and low-cost solar thermal distiller
with multi-layered distillation stages, each of which is made of two opposing hydro-
philic layers (as evaporator and condenser, respectively) separated by a hydropho-
bic microporous membrane. Latent heat of vaporization can be reused multiple
times passively without mechanical ancillaries before it is lost to the environment
in such an N-stage device, contributing to a distillate flow rate of almost 3 L m 2
h 1 from seawater at less than 1 sun.

Besides, solar-driven zero liquid discharge (ZLD) desalination from waste brine water
has recently emerged as a new important application of solar evaporation.154,155
Compared with conventional ZLD desalination technologies, solar-driven ZLD
desalination produces solid salt as the only byproduct and uses sunlight as the
only energy source, making it less energy intensive and more cost-effective. Mi
and co-workers154 demonstrated a mangrove-inspired solar evaporator based on
a synthetic GO leaf with two kinds of configurations: floating on water and lifted
above water with a tree-like configuration. The tree-like synthetic GO leaf delivered
a higher energy efficiency than the floating configuration (78% versus 54%) with an

706 Joule 3, 683–718, March 20, 2019


Joule 3, 683–718, March 20, 2019 707
Figure 16. Hybrid Systems for Solar Desalination and Electric Generation Induced by Salinity Gradients, Steam Evaporation, and Triboelectric
Nanogenerators (TENGs)
(A–C) A salinity-driven hybrid system: (A) structure of the hybrid device, (B) mechanism of electricity generation from the salinity difference, and (C)
photograph of a hybrid device. Reproduced from Zhou et al., 13 with permission.
(D–F) Steam-driven hybrid system: (D) graphical illustration of the self-contained carbon sponge (CS), (E) schematic diagram of steam generation-
induced electric potential by the CS, and (F) piezo-pyroelectric output voltages of the PVDF films. Reproduced from Ho et al., 157 with permission.
(G–L) A TENG-driven hybrid system. (G) Graphical illustration of the integral prototype for condensate collection and triboelectric energy generation,
along with a diagram of a TENG with water flowing down and swinging in the round bottom vessel. (H) The condensate collected under focused sunlight
after the first 60 min; open-circuit voltage and closed circuit current measurements of the TENG device with flowing (I and J) and swinging (K and L)
configurations. Reproduced from Ho et al., 160 with permission.

evaporation rate of 2 L m 2 h 1. Moreover, the GO leaf demonstrated relatively sta-


ble performance despite gradual and eventually severe accumulation of salt crystals
on the leaf surface during a long-term evaporation experiment with a 15 wt % NaCl
solution, representing a promising step toward stable and high-efficiency ZLD
desalination.

In a more recent study, Wang’s group155 demonstrated a smart design concept of a


solar-driven ZLD desalination device using a 3D cup structure in which the salt pre-
cipitation surface (outer surface of the 3D cup evaporator) is separated from the
evaporative surface (bottom and inner surface of the 3D cup evaporator). In this
case, continuous salt precipitation will not affect the evaporation process, even at
an extremely high salt concentration of 25 wt %, resulting in stable device operation
for more than 120 h. This new solar evaporator design opens opportunities for
greener ZLD desalination, especially for high-salinity brine treatment in desalination
plants as well as hypersaline industrial wastewater treatment.

Electricity Generation
From the point of view of energy conversion, the solar evaporation process is also an
efficient energy harvesting method, which converts solar energy into thermal en-
ergy, stored in forms of hot vapor or water. However, there is huge energy waste be-
tween the solar energy input and the final cold water we receive. Rational utilization
or saving the energy during the solar evaporation process will bring more opportu-
nities in dealing with both water and energy scarcities. Recently, Zhou’s group13
proposed a hybrid system that can generate 1 W m 2 of electricity power while
simultaneously demonstrating a high evaporation efficiency (Figures 16A–16C).
The extra power comes from the evaporation-induced salinity gradient across the
evaporator, which the authors claim can theoretically generate 12.5 W m 2 real-
time salinity power between the surface and the bulk solution during solar evapora-
tion under 1 sun irradiance. But because none of the ions at the concentrated area
were passing through the Nafion, the real electricity power recovered from the
salinity power was actually less than 10%. In another work,156 naturally generated
vapor was demonstrated to be capable of generating electricity with a high voltage
of 1 V when it transports through the surface of an amorphous carbon layer. One un-
desired limitation of this design is that the demonstrated device fails to collect clean
water simultaneously, resulting in a significant waste of energy.

Ho and co-workers157 successfully addressed this issue by combining solar-driven


distillation with electricity generation using a self-contained monolithic carbon
sponge evaporator design (Figures 16D–16F). Thermal energy in the hot vapor
can also be harvested by pyroelectric nanogenerators,158 thermoelectric mod-
ules,159 and triboelectric nanogenerators.160 For example, Ho’s group160 proposed
another hybrid system (Figures 16G–16L) where the energy released from the
condensation process was utilized to generate electricity via an integral triboelectric

708 Joule 3, 683–718, March 20, 2019


nanogenerator (TENG). In the flowing configuration, triboelectric signals can be
generated from the electrification of condensate water and polytetrafluoroethylene
(PTFE) film (Figures 16I and 16S). In the swinging configuration, the round-bottom
prototype structure is able to harness omnidirectional mechanical energy (e.g.,
wind energy), which causes it to swing in various directions. The swinging motion
in turn triggers movement of the collected condensate inside the round bottom to
generate triboelectric signals (Figures 16K and 16L).

In a more recent study, Zhu’s group159 developed an integrated interfacial solar


evaporation-thermoelectric module system to simultaneously produce clean water
and electricity using contaminated water as a source material and solar as the only
input energy. Compared with the traditional steam condensing process to form
clean water, in which a large amount of latent heat energy is wasted, the inte-
grated system is designed to store and recycle steam enthalpy as thermal energy,
which can be further converted to electricity via a thermoelectric module (Figures
17A and 17B). The authors use a scalable and low-cost nonwoven material coated
with graphite particles as a highly efficient yet flexible and foldable light absorber,
which imparts the developed integrated system with the potential for practical
large-scale applications (Figure 17C). The highly absorptive light absorber along
with good thermal insulation of the interfacial heating system contribute to high
evaporation rates and solar-to-steam efficiencies (up to 81.7%) at various solar
densities (Figure 17D). The collected water from steam condensation is of high pu-
rity, meeting the WHO standards for potable water (Figure 17E). Thermal energy
gained from the hot vapor is stored in the thermal storage area with the chamber
temperature maintained at 100 C and room temperature (25 C) at the two sides.
Under this temperature gradient, the coupled thermoelectric module is able to
convert thermal energy into electricity with a maximum open-circuit voltage of
3.87 V and short-circuit current of 0.55 A at 30 kW m 2 (Figures 17F and 17G),
which can be further improved to 4.17 V and 0.61 A (corresponding to a maximum
efficiency of 1.23%) if a superheated steam technique is used to increase the steam
temperature. The generated electricity can drive the continuous operation of an
electric fan (1 W) and 28 light-emitting diodes (total power 1.5 W), as shown in
Figure 17H.

This exciting progress of combining solar evaporation with electricity generation


made in the past few years provides this technology with greater prospects and po-
tential for addressing challenges in global water and energy scarcity. We believe this
kind of multifunctional device, given its sustainable nature and ability to simulta-
neously address fresh water and energy needs, will receive more attention in the
foreseeable future. However, continuous efforts in material and structural design
and interdisciplinary integration are still needed to further increase the energy con-
version efficiency, as well as lower material and device costs toward commercial
applications.

Other Applications
Solar evaporation also demonstrates great potential for converting solar energy to
other forms of energy, such as chemical energy (via fuel production) and mechanical
energy.15,161,162 In a study led by Grimes and co-workers,15 nitrogen-doped
titanium dioxide nanotube arrays were used to photocatalytically convert CO2 and
water vapor into hydrocarbon fuels under the assistance of solar irradiation (Fig-
ure 18A). Nitrogen doping was used to narrow the band gap of titanium dioxide
nanotube for enhanced light absorption, while ultrathin tube thickness was fabri-
cated to enhance the loading of catalyst nanoparticles. Such a device can achieve

Joule 3, 683–718, March 20, 2019 709


Figure 17. Hybrid System for Water Purification and Electric Generation via Storing and Recycling the Steam Enthalpy
(A) Traditional steam condenses to form water with enthalpy lost to the environment.
(B) Storage and recycling of interfacial solar steam enthalpy for simultaneous generation of clean water and electricity.
(C) Photographs and SEM images of nonwoven and graphite/nonwoven films.
(D) Evaporation rates and efficiencies of the graphite/nonwoven film under various power densities.
(E) Concentrations of various metal ions before and after treatment (the blue dashed line refers to the WHO standards for drinking water).
(F and G) Open-circuit voltage (F) and short-circuit current (G) over time under different solar irradiation.
(H) Photographs of an electric fan and light-emitting diode powered by interfacial solar evaporation. Reproduced from Zhu et al., 159 with permission.

a hydrocarbon production rate of 111 ppm cm 2 h 1, or z160 mL/(g h) under


ambient solar flux, which is at least 20 times higher than in previously published re-
ports. Recently, Halas et al.161 demonstrated a smart strategy to produce cellulosic
bioethanol fully off grid by combining solar steam processing and solar distillation.
The nanoparticle-enabled solar-generated steam was utilized to successfully hydro-
lyze feedstock into sugars, while solar steam distillation was chosen as the final pro-
cessing step to purify ethanol. This entirely off-grid solar-assisted production
method demonstrates great potential in realizing the long-dreamed-of goal of sus-
tainable and greener production of cellulosic bioethanol and other chemical fuels.

Besides water and energy generation, solar evaporation also shows great potential
in response to global public medical and health needs. Currently, there are still over
650 million people around the world who lack access to safe water, and one-quarter
of the world’s population lacks access to electricity.14,163 The high-temperature

710 Joule 3, 683–718, March 20, 2019


Figure 18. Other Potential Applications of Solar Evaporation
(A) Solar evaporation-enabled fuel production. Reproduced from Grimes et al., 15 with permission. Copyright 2009 American Chemical Society.
(B) Solar steam sterilization. Reproduced from Oara et al., 14 with permission.

steam formed by solar evaporation therefore provides a good option for sterilization
in areas where electric sterilization systems are not available.14,164,165 Oara et al.14
have reported a compact solar sterilization system based on the broadband light ab-
sorption of nanoparticles (Figure 18B). By using a 44-in solar dish concentrator, the
solar autoclave can release and maintain high-temperature steam (132 C) for 5 min,
which can provide the sanitation requirements for a 14.2 L volume content. In
another example, Deng and co-workers164 demonstrated a solar-driven sterilization
device based on interfacial solar evaporation using a floating rGO-based black
membrane as a light absorber, which enables energy efficient steam sterilization
(with a steam temperature of >120 C). The capital cost for such a solar-driven ster-
ilization device is significantly lower than commercial autoclaves because of the use
of only low-cost materials without complex facilities, which offers a potential cost-
effective solution to meet the need for sterilization in undeveloped areas that lack
electrical power but have ample solar radiation. With the fast growth of both novel
materials and structures, and substantial enhancement in evaporation performance
in the past few years, we believe that revisiting the solar steam sterilization technique
is appropriate.165

Challenges and Future Perspective


Great achievements have been made in the past decade in the development of pho-
tothermal and substrate materials, as well as diverse structural engineering strate-
gies involving light absorption, water transportation, thermal insulation, and interfa-
cial properties. While there are emerging applications that take advantage of this
intriguing technology, there remain several challenges that require further study
to bring solar evaporation into practice for the benefit of daily society. Both funda-
mental research and practical explorations are warranted to make this clean, green,
and sustainable technology possible.

Joule 3, 683–718, March 20, 2019 711


Standardized Measurements and Evaluation
A long-standing challenge for evaluating the performance of solar evaporation de-
vices is the lack of standardized measurement techniques and calculation of evapo-
ration efficiency. As a result, it is difficult and unreliable to directly compare reported
results, which has been an obstacle to the healthy development of this field. One
possible solution is to clearly state the specific calculation formula and how these pa-
rameters are determined in detail. In addition, the natural evaporation contribution
needs to be subtracted from the final evaporation rate calculation to eliminate the
influence of environment, such as wind and temperature. It is also desirable for
the solar evaporation community to establish widely accepted guidelines as stan-
dards for measurements and calculations.

Multi-level Stability against Water, Heat, Salt, Bacteria, and Weather during Long-
Term Operation
The long-term stability of photothermal materials and structures when dealing with
real water sources (e.g., seawater, river water, groundwater, industrial contaminated
water, and city sewage) has been a significant challenge. Research interests in solar
evaporation have mainly focused on achieving higher evaporation efficiencies, while
limited attention has been paid to the stability of the materials and device perfor-
mance against water, heat, salt, bacteria, and weather for long-term operation. Sta-
bility is of great importance, especially for practical applications. More efforts are
needed to investigate and improve this aspect through either material engineering,
such as the use of protective coatings, surface chemistry, and hybridization, or
structural designs with additional functions to protect the device, such as by pore en-
gineering and structure integration. For example, clogging of salt is considered a
major issue during long-term desalination, especially for high-salinity water sources.
Surface wettability and pore structure engineering has proven to be effective in miti-
gating the clogging issue, yet there is a trade-off with evaporation efficiency.
Resolving this issue remains an open question for future study.

Special attention needs to be paid to biomass-based materials to improve their sta-


bility against water, heat, and bacteria. For plasmonic nanoparticle systems, the
chemical stability of the highly active nanoparticles also needs to be improved by
coating, surface functionalization, or surface energy engineering. Some contami-
nated water sources containing volatile organic components also face another chal-
lenge in that organic compounds tend to evaporate along with the collected water,
leading to an impure water product. Structural and material designs that can surpass
the evaporation of organic components without affecting the water evaporation is
highly desired. Another potential strategy is to separate the organic component
evaporation process from the water evaporation process, thus enabling the collec-
tion of clean water separately.

Scalable and Cost-Effective Production toward Practical Applications


For practical applications, scalability and cost are two key factors that will determine
the competitiveness of this technology. From a product point of view, large-scale
manufacturing and low cost are desirable and beneficial to promote acceptance
by potential customers. Previous and current studies of solar evaporation have
generally failed to provide the estimated manufacturing scalability and cost of the
developed materials and/or structures, which creates difficulty in evaluating the
technology’s potential for practical applications. A greater focus on these areas
when presenting novel research findings would benefit the field. Additionally, stra-
tegies for developing new materials and structures with potentially lower price and/
or larger scalability warrant further study.

712 Joule 3, 683–718, March 20, 2019


Figure 19. Schematic Illustration of a Spider
Chart of Future Solar Evaporation Material and
Structural Designs
Future solar evaporation material and structural
designs must balance efficiency, cost,
scalability, stability, and adaptability for the
development of commercially viable solar
thermal systems.

Further Understanding the Structure-Property Relationships of Solar Evaporation


Materials and Structures
Although many new materials and structures have been developed in the past
decade, we often lack a fundamental understanding of the dynamic kinetics of these
processes and the underlying structure-property relationships. From a material and
structural point of view, better understanding of the light transfer, water transport,
and thermal diffusion behaviors and kinetics, and the corresponding structure-prop-
erty relationships, is critical for future solar evaporator designs and performance
optimization. For example, water transport behavior is highly related to the pore
structure (e.g., porosity, pore size, and tortuosity) and surface properties (e.g.,
wettability). How the pore structure, particularly pore size and tortuosity, and surface
properties influence water transport behavior is yet to be fully explored. Thermal
diffusion behavior, including conduction, radiation, and convection, is another
important aspect that needs more fundamental understanding. Heat localization,
thermal concentration, and cold vapor generation represent three smart strategies
for efficient thermal utilization. Surpassing thermal loss by minimizing thermal con-
duction has been the research focus in the past few years, while very limited atten-
tion has been paid to the convection and radiation parts. Gaining energy from the
environment by cold vapor generation and taking advantage of newly emerging
daytime cooling technologies suggest some new directions to further enhance the
thermal efficiency toward surpassing thermal convection and/or radiation.

In terms of fundamental understanding, simulation can be a powerful tool when


combined with experimental observations to gain insights into the fundamental be-
haviors and kinetics of light, water, vapor, and thermal-related processes. Addition-
ally, real-time monitoring of these processes via in situ characterization technologies
is also desirable for gaining fundamental insights of the underlying mechanisms.

Exploring New Materials and Structural Design


The past decade has witnessed tremendous progress in new material and structural
designs for solar evaporation, which have increased the solar evaporation efficiency
to a record-high level of >90% under ambient solar flux. We anticipate that new
material and structural designs will further enhance the solar evaporation perfor-
mance in various aspects that are not limited to evaporation efficiency over the
next 5–10 years. Knowledge in smart material and structural design is needed to
further take this technology to the next development level. Nanotechnology will
continue to play a key role in this process by providing diverse and powerful capa-
bilities in new material and structural design and engineering. Hybridization, on

Joule 3, 683–718, March 20, 2019 713


both material and structural levels, will be another powerful strategy to enable new
features, functions, or enhanced performance. In terms of structural design, nature is
a good teacher. Learning from nature to mimic the structure and behaviors in plants,
animals, or other natural phenomena represents a promising direction for future
solar evaporator designs.

Exploring New Functions and Applications


Since its emergence, solar evaporation has been used for diverse applications,
including water purification, desalination, electric and power generation, photoca-
talysis, solar-chemical fuel production, and sterilization. However, these applica-
tions are still in their infancy, and many further functions and applications need future
exploration. Instead of pursuing only higher thermal efficiency, the research commu-
nity also needs to explore how to use the generated vapor for clean water, power,
and fuel generation in more practical ways. Integrating new functions into state-
of-the-art solar evaporation systems to further broaden the technology’s applica-
tions will be the research focus in the next decade. For example, zero liquid
discharge systems are necessary to meet increasing regulations on direct brine
discharge. As an emerging application, solar-evaporation-driven zero liquid
discharge desalination offers a promising solution toward sustainable and cost-
effective brine or hypersaline water treatment. Interdisciplinary convergence of
this technology with other emerging fields, such as nanofluidics, thermoelectricity,
rechargeable batteries, photo- and electro-catalysis, organic fuel production, radia-
tive cooling, etc., will bring this technology to a broader prospect and future.

Given the abundant, renewable, and widely distributed resources from solar energy
and diverse water sources, this green process with zero carbon footprint and facile
operation without complex facilities makes solar evaporation one of the most prom-
ising technologies toward clean water, energy, and fuel generation and conversion.
Despite its multiple challenges in various aspects, we anticipate the next decade will
witness the boom of this green technology in a variety of application fields, playing a
critical role in addressing global crises, especially the pressing global water short-
ages and growing demand for clean energy. With considerable efforts to bridge
the gaps between state-of-the-art solar evaporation systems and future scalable
and cost-effective practical applications, this technology will be able to balance ef-
ficiency, cost, scalability, stability, and adaptability in the years to come (Figure 19).

ACKNOWLEDGMENTS
We acknowledge the support of the Maryland NanoCenter and its AIMLab.

AUTHOR CONTRIBUTIONS
L.H. and C.C. proposed the topic of this review. L.H., C.C., and Y.K. investigated the
literature and co-wrote the manuscript.

REFERENCES
1. Crabtree, G.W., and Lewis, N.S. (2007). Solar 4. Oki, T., and Kanae, S. (2006). Global water production by sunlight. Environ. Sci.
energy conversion. Phys. Today 60, 37–42. hydrological cycles and world water Nano 5, 1078–1089.
resources. Science 313, 1068–1072.
2. Cavusoglu, A.H., Chen, X., Gentine, P., and 7. Gao, M., Zhu, L., Peh, C.K.N., and Ho, G.W.
Sahin, O. (2017). Potential for natural 5. Qiblawey, H.M., and Banat, F. (2008). Solar (2018). Solar absorber material and system
evaporation as a reliable renewable energy thermal desalination technologies. designs for photothermal water vaporization
resource. Nat. Commun. 8, 617. Desalination 220, 633–644. towards clean water and energy production.
Energy Environ. Sci. https://doi.org/10.1039/
3. Lewis, N.S. (2016). Research opportunities to 6. Wang, P. (2018). Emerging investigator series: C8EE01146J.
advance solar energy utilization. Science 351, the rise of nano-enabled photothermal
8. Chandrashekara, M., and Yadav, A. (2017).
aad1920. materials for water evaporation and clean
Water desalination system using solar heat: a

714 Joule 3, 683–718, March 20, 2019


review. Renew. Sustain. Energy Rev. 67, 1308– 23. Liu, X., Wang, X., Huang, J., Cheng, G., and 38. Liu, Z., Song, H., Ji, D., Li, C., Cheney, A., Liu,
1330. He, Y. (2018). Volumetric solar steam Y., Zhang, N., Zeng, X., Chen, B., Gao, J., et al.
generation enhanced by reduced graphene (2017). Extremely cost-effective and efficient
9. Khawaji, A.D., Kutubkhanah, I.K., and Wie, oxide nanofluid. Appl. Energy 220, 302–312. solar vapor generation under
J.-M. (2008). Advances in seawater nonconcentrated illumination using thermally
desalination technologies. Desalination 221, 24. Yu, F., Chen, Y., Liang, X., Xu, J., Lee, C., isolated black paper. Global Chall. 1,
47–69. Liang, Q., Tao, P., and Deng, T. (2017). 1600003.
Dispersion stability of thermal nanofluids.
10. Elimelech, M., and Phillip, W.A. (2011). The Progress in Natural Science: Materials 39. Li, X., Xu, W., Tang, M., Zhou, L., Zhu, B., Zhu,
future of seawater desalination: energy, International 27, 531–542. S., and Zhu, J. (2016). Graphene oxide-based
technology, and the environment. Science efficient and scalable solar desalination under
333, 712–717. 25. Prasher, R., Phelan, P.E., and Bhattacharya, P. one sun with a confined 2D water path. Proc.
(2006). Effect of aggregation kinetics on the Natl. Acad. Sci. USA 113, 13953–13958.
11. Zhou, L., Tan, Y., Wang, J., Xu, W., Yuan, Y., thermal conductivity of nanoscale colloidal
Cai, W., Zhu, S., and Zhu, J. (2016). 3D self- solutions (nanofluid). Nano Lett. 6, 1529–1534. 40. Zhu, M., Li, Y., Chen, G., Jiang, F., Yang, Z.,
assembly of aluminium nanoparticles for Luo, X., Wang, Y., Lacey, S.D., Dai, J., Wang,
plasmon-enhanced solar desalination. Nat. 26. Jin, H., Lin, G., Bai, L., Zeiny, A., and Wen, D. C., et al. (2017). Tree-inspired design for high-
Photonics 10, 393–398. (2016). Steam generation in a nanoparticle- efficiency water extraction. Adv. Mater. 29,
based solar receiver. Nano Energy 28, 1704107.
12. Higgins, M.W., Shakeel Rahmaan, A.S., 397–406.
Devarapalli, R.R., Shelke, M.V., and Jha, N. 41. Xu, W., Hu, X., Zhuang, S., Wang, Y., Li, X.,
(2018). Carbon fabric based solar steam 27. Zeng, Y., Yao, J., Horri, B.A., Wang, K., Wu, Y., Zhou, L., Zhu, S., and Zhu, J. (2018). Flexible
generation for waste water treatment. Sol. Li, D., and Wang, H. (2011). Solar evaporation and salt resistant Janus absorbers by
Energy 159, 800–810. enhancement using floating light-absorbing electrospinning for stable and efficient solar
magnetic particles. Energy Environ. Sci. 4, desalination. Adv. Energy Mater. 8, 1702884.
13. Yang, P., Liu, K., Chen, Q., Li, J., Duan, J., Xue, 4074–4078.
G., Xu, Z., Xie, W., and Zhou, J. (2017). Solar- 42. Cao, F., McEnaney, K., Chen, G., and Ren, Z.
driven simultaneous steam production and 28. Wang, Z., Liu, Y., Tao, P., Shen, Q., Yi, N., (2014). A review of cermet-based spectrally
electricity generation from salinity. Energy Zhang, F., Liu, Q., Song, C., Zhang, D., Shang, selective solar absorbers. Energy Environ. Sci.
Environ. Sci. 10, 1923–1927. W., et al. (2014). Bio-inspired evaporation 7, 1615–1627.
through plasmonic film of nanoparticles at the
14. Neumann, O., Feronti, C., Neumann, A.D., air-water interface. Small 10, 3234–3239. 43. Lim, D.K., Barhoumi, A., Wylie, R.G., Reznor,
Dong, A., Schell, K., Lu, B., Kim, E., Quinn, M., G., Langer, R.S., and Kohane, D.S. (2013).
Thompson, S., Grady, N., et al. (2013). 29. Ito, Y., Tanabe, Y., Han, J., Fujita, T., Tanigaki, Enhanced photothermal effect of plasmonic
Compact solar autoclave based on steam K., and Chen, M. (2015). Multifunctional nanoparticles coated with reduced graphene
generation using broadband light-harvesting porous graphene for high-efficiency steam oxide. Nano Lett. 13, 4075–4079.
nanoparticles. Proc. Natl. Acad. Sci. USA 110, generation by heat localization. Adv. Mater.
11677–11681. 27, 4302–4307. 44. Chen, Y., Wang, L., and Shi, J. (2016). Two-
dimensional non-carbonaceous materials-
15. Varghese, O.K., Paulose, M., LaTempa, T.J., 30. Ghasemi, H., Ni, G., Marconnet, A.M., enabled efficient photothermal cancer
and Grimes, C.A. (2009). High-rate solar Loomis, J., Yerci, S., Miljkovic, N., and Chen, therapy. Nano Today 11, 292–308.
photocatalytic conversion of CO2 and water G. (2014). Solar steam generation by heat
vapor to hydrocarbon fuels. Nano Lett. 9, localization. Nat. Commun. 5, 4449. 45. Yang, M.Q., Gao, M., Hong, M., and Ho, G.W.
731–737. (2018). Visible-to-NIR photon harvesting:
31. Jiang, Q., Tian, L., Liu, K.K., Tadepalli, S., progressive engineering of catalysts for solar-
16. Kaushal, A., and Varun. (2010). Solar stills: a Raliya, R., Biswas, P., Naik, R.R., and powered environmental purification and fuel
review. Renew. Sustain. Energy Rev. 14, Singamaneni, S. (2016). Bilayered biofoam for production. Adv. Mater. 30, 1802894.
446–453. highly efficient solar steam generation. Adv.
Mater. 28, 9400–9407. 46. Govorov, A.O., and Richardson, H.H. (2007).
17. Kabeel, A.E., and El-Agouz, S.A. (2011). Generating heat with metal nanoparticles.
Review of researches and developments on 32. Shang, W., and Deng, T. (2016). Solar steam Nano Today 2, 30–38.
solar stills. Desalination 276, 1–12. generation: steam by thermal concentration.
Nat. Energy 1, 16133. 47. Brongersma, M.L., Halas, N.J., and
18. Neumann, O., Urban, A.S., Day, J., Lal, S., Nordlander, P. (2015). Plasmon-induced hot
Nordlander, P., and Halas, N.J. (2013). Solar 33. Tao, P., Ni, G., Song, C., Shang, W., Wu, J., carrier science and technology. Nat.
vapor generation enabled by nanoparticles. Zhu, J., Chen, G., and Deng, T. (2018). Solar- Nanotechnol. 10, 25–34.
ACS Nano 7, 42–49. driven interfacial evaporation. Nat. Energy 3,
1031–1041. 48. Liu, T., and Li, Y. (2016). Photocatalysis:
19. Zhao, D., Duan, H., Yu, S., Zhang, Y., He, J., plasmonic solar desalination. Nat. Photonics
Quan, X., Tao, P., Shang, W., Wu, J., Song, C., 34. Gao, T., Li, Y., Chen, C., Yang, Z., Kuang, Y., 10, 361–362.
et al. (2015). Enhancing localized evaporation Jia, C., Song, J., Hitz, E.M., Liu, B., Huang, H.,
through separated light absorbing centers et al. (2018). Architecting a floatable, durable, 49. Gao, M., Connor, P.K.N., and Ho, G.W. (2016).
and scattering centers. Sci. Rep. 5, 17276. and scalable steam generator: hydrophobic/ Plasmonic photothermic directed broadband
hydrophilic bifunctional structure for solar sunlight harnessing for seawater catalysis and
20. Ni, G., Miljkovic, N., Ghasemi, H., Huang, X., evaporation enhancement. Small Methods desalination. Energy Environ. Sci. 9, 3151–
Boriskina, S.V., Lin, C.-T., Wang, J., Xu, Y., 2018, 1800176. 3160.
Rahman, M.M., Zhang, T., et al. (2015).
Volumetric solar heating of nanofluids for 35. Liu, G., Xu, J., and Wang, K. (2017). Solar water 50. Berciaud, S., Cognet, L., and Lounis, B. (2005).
direct vapor generation. Nano Energy 17, evaporation by black photothermal sheets. Photothermal absorption spectroscopy of
290–301. Nano Energy 41, 269–284. individual semiconductor nanocrystals. Nano
Lett. 5, 2160–2163.
21. Hogan, N.J., Urban, A.S., Ayala-Orozco, C., 36. Zhu, L., Gao, M., Peh, C.K.N., and Ho, G.W.
Pimpinelli, A., Nordlander, P., and Halas, N.J. (2018). Solar-driven photothermal 51. Song, Y., Cretin, B., Todorovic, D.M., and
(2014). Nanoparticles heat through light nanostructured materials designs and Vairac, P. (2008). Study of photothermal
localization. Nano Lett. 14, 4640–4645. prerequisites for evaporation and catalysis vibrations of semiconductor cantilevers
applications. Mater. Horiz. 5, 323–343. near the resonant frequency. J. Phys. D Appl.
22. Fang, Z., Zhen, Y.R., Neumann, O., Polman, Phys. 41.
A., Garcı́a de Abajo, F.J., Nordlander, P., and 37. Deng, Z., Zhou, J., Miao, L., Liu, C., Peng, Y.,
Halas, N.J. (2013). Evolution of light-induced Sun, L., and Tanemura, S. (2017). The 52. Liu, H., Chen, C., Wen, H., Guo, R., Williams,
vapor generation at a liquid-immersed emergence of solar thermal utilization: solar- N.A., Wang, B., Chen, F., and Hu, L. (2018).
metallic nanoparticle. Nano Lett. 13, 1736– driven steam generation. J. Mater. Chem. A 5, Narrow bandgap semiconductor decorated
1742. 7691–7709. wood membrane for high-efficiency

Joule 3, 683–718, March 20, 2019 715


solar-assisted water purification. J. Mater. solar steam generation. Sol. Energy 157, 79. Cui, L., Zhang, P., Xiao, Y., Liang, Y., Liang, H.,
Chem. A 6, 18839–18846. 35–46. Cheng, Z., and Qu, L. (2018). High rate
production of clean water based on the
53. Wang, J., Li, Y., Deng, L., Wei, N., Weng, Y., 66. Zhu, M., Li, Y., Chen, F., Zhu, X., Dai, J., Li, Y., combined photo-electro-thermal effect of
Dong, S., Qi, D., Qiu, J., Chen, X., and Wu, T. Yang, Z., Yan, X., Song, J., Wang, Y., et al. graphene architecture. Adv. Mater. 30,
(2017). High-performance photothermal (2018). Plasmonic wood for high-efficiency e1706805.
conversion of narrow-bandgap Ti2O3 solar steam generation. Adv. Energy Mater. 8,
nanoparticles. Adv. Mater. 29, 1603730. 1701028. 80. Fu, Y., Wang, G., Ming, X., Liu, X., Hou, B.,
Mei, T., Li, J., Wang, J., and Wang, X. (2018).
54. Murakami, T., Nakatsuji, H., Inada, M., 67. Liu, Y., Yu, S., Feng, R., Bernard, A., Liu, Y., Oxygen plasma treated graphene aerogel as
Matoba, Y., Umeyama, T., Tsujimoto, M., Zhang, Y., Duan, H., Shang, W., Tao, P., Song, a solar absorber for rapid and efficient solar
Isoda, S., Hashida, M., and Imahori, H. (2012). C., et al. (2015). A bioinspired, reusable, steam generation. Carbon 130, 250–256.
Photodynamic and photothermal effects of paper-based system for high-performance
semiconducting and metallic-enriched single- large-scale evaporation. Adv. Mater. 27, 81. Hao, W., Chiou, K., Qiao, Y., Liu, Y., Song, C.,
walled carbon nanotubes. J. Am. Chem. Soc. 2768–2774. Deng, T., and Huang, J. (2018). Crumpled
134, 17862–17865. graphene ball-based broadband solar
68. Sun, W., Zhong, G., Kübel, C., Jelle, A.A., absorbers. Nanoscale 10, 6306–6312.
55. Wu, M.C., Deokar, A.R., Liao, J.H., Shih, P.Y., Qian, C., Wang, L., Ebrahimi, M., Reyes, L.M.,
and Ling, Y.C. (2013). Graphene-based 82. Kim, K., Yu, S., An, C., Kim, S.W., and Jang,
Helmy, A.S., and Ozin, G.A. (2017). Size-
photothermal agent for rapid and effective J.H. (2018). Mesoporous three-dimensional
tunable photothermal germanium
killing of bacteria. ACS Nano 7, 1281–1290. graphene networks for highly efficient solar
nanocrystals. Angew. Chem. 56, 6329–6334.
desalination under 1 sun illumination. ACS
56. Zhang, L., Tang, B., Wu, J., Li, R., and Wang, P. Appl. Mater. Interfaces 10, 15602–15608.
69. Chen, M., He, Y., Huang, J., and Zhu, J. (2016).
(2015). Hydrophobic light-to-heat conversion Synthesis and solar photo-thermal conversion
membranes with self-healing ability for 83. Yang, Y., Zhao, R., Zhang, T., Zhao, K., Xiao, P.,
of Au, Ag, and Au-Ag blended plasmonic Ma, Y., Ajayan, P.M., Shi, G., and Chen, Y.
interfacial solar heating. Adv. Mater. 27, 4889– nanoparticles. Energy Convers. Manage. 127,
4894. (2018). Graphene-based standalone solar
293–300. energy converter for water desalination and
57. Zhao, F., Zhou, X., Shi, Y., Qian, X., Alexander, purification. ACS Nano 12, 829–835.
70. Almond, D.P., Patel, P., and Patel, P. (1996).
M., Zhao, X., Mendez, S., Yang, R., Qu, L., and Photothermal Science and Techniques 84. Zhang, P., Liao, Q., Zhang, T., Cheng, H.,
Yu, G. (2018). Highly efficient solar vapour (Springer). Huang, Y., Yang, C., Li, C., Jiang, L., and Qu, L.
generation via hierarchically nanostructured
(2018). High throughput of clean water
gels. Nat. Nanotechnol. 13, 489–495. 71. Zhu, G., Xu, J., Zhao, W., and Huang, F. (2016). excluding ions, organic media, and bacteria
Constructing black titania with unique from defect-abundant graphene aerogel
58. Politano, A., Argurio, P., Di Profio, G., Sanna,
nanocage structure for solar desalination. under sunlight. Nano Energy 46, 415–422.
V., Cupolillo, A., Chakraborty, S., Arafat, H.A.,
ACS Appl. Mater. Interfaces 8, 31716–31721.
and Curcio, E. (2017). Photothermal 85. Shi, L., Wang, Y., Zhang, L., and Wang, P.
membrane distillation for seawater 72. Ding, D., Huang, W., Song, C., Yan, M., Guo, (2017). Rational design of a bi-layered
desalination. Adv. Mater. 29, 1603504. C., and Liu, S. (2017). Non-stoichiometric reduced graphene oxide film on polystyrene
MoO3-x quantum dots as a light-harvesting foam for solar-driven interfacial water
59. Jain, P.K., Huang, X., El-Sayed, I.H., and El-
Sayed, M.A. (2008). Noble metals on the
material for interfacial water evaporation. evaporation. J. Mater. Chem. A 5, 16212–
Chem. Commun. (Camb.) 53, 6744–6747. 16219.
nanoscale: optical and photothermal
properties and some applications in imaging, 86. Hu, X., Xu, W., Zhou, L., Tan, Y., Wang, Y., Zhu,
73. Chen, R., Wu, Z., Zhang, T., Yu, T., and Ye, M.
sensing, biology, and medicine. Acc. Chem. S., and Zhu, J. (2017). Tailoring graphene
(2017). Magnetically recyclable self-
Res. 41, 1578–1586. oxide-based aerogels for efficient solar steam
assembled thin films for highly efficient water
evaporation by interfacial solar heating. RSC generation under one sun. Adv. Mater. 29,
60. Zhou, L., Tan, Y., Ji, D., Zhu, B., Zhang, P., Xu,
J., Gan, Q., Yu, Z., and Zhu, J. (2016). Self- Adv. 7, 19849–19855. 1604031.
assembly of highly efficient, broadband 87. Li, Y., Gao, T., Yang, Z., Chen, C., Kuang, Y.,
74. Shi, Y., Li, R., Jin, Y., Zhuo, S., Shi, L., Chang, J.,
plasmonic absorbers for solar steam Song, J., Jia, C., Hitz, E.M., Yang, B., and Hu,
Hong, S., Ng, K.-C., and Wang, P. (2018). A 3D
generation. Sci. Adv. 2, e1501227. L. (2017). Graphene oxide-based evaporator
photothermal structure toward improved
energy efficiency in solar steam generation. with one-dimensional water transport
61. Fu, Y., Mei, T., Wang, G., Guo, A., Dai, G., enabling high-efficiency solar desalination.
Wang, S., Wang, J., Li, J., and Wang, X. (2017). Joule 2, 1171–1186.
Nano Energy 41, 201–209.
Investigation on enhancing effects of Au
nanoparticles on solar steam generation in 75. Vélez-Cordero, J.R., and Hernández-Cordero,
88. Liu, K.K., Jiang, Q., Tadepalli, S., Raliya, R.,
graphene oxide nanofluids. Appl. Therm. J. (2015). Heat generation and conduction in
Biswas, P., Naik, R.R., and Singamaneni, S.
Eng. 114, 961–968. PDMS-carbon nanoparticle membranes
(2017). Wood-graphene oxide composite for
irradiated with optical fibers. Int. J. Therm. Sci.
highly efficient solar steam generation and
62. Bae, K., Kang, G., Cho, S.K., Park, W., Kim, K., 96, 12–22.
desalination. ACS Appl. Mater. Interfaces 9,
and Padilla, W.J. (2015). Flexible thin-film 7675–7681.
black gold membranes with ultrabroadband 76. Jia, C., Li, Y., Yang, Z., Chen, G., Yao, Y., Jiang,
plasmonic nanofocusing for efficient solar F., Kuang, Y., Pastel, G., Xie, H., Yang, B., et al. 89. Wang, G., Fu, Y., Guo, A., Mei, T., Wang, J., Li,
vapour generation. Nat. Commun. 6, 10103. (2017). Rich mesostructures derived from J., and Wang, X. (2017). Reduced graphene
natural woods for solar steam generation. oxide-polyurethane nanocomposite foam as
63. Zhou, L., Zhuang, S., He, C., Tan, Y., Wang, Z., Joule 1, 588–599. a reusable photoreceiver for efficient solar
and Zhu, J. (2017). Self-assembled spectrum steam generation. Chem. Mater. 29, 5629–
selective plasmonic absorbers with tunable 77. Yang, J., Pang, Y., Huang, W., Shaw, S.K., 5635.
bandwidth for solar energy conversion. Nano Schiffbauer, J., Pillers, M.A., Mu, X., Luo, S.,
Energy 32, 195–200. Zhang, T., Huang, Y., et al. (2017). 90. Chen, T., Wang, S., Wu, Z., Wang, X., Peng, J.,
Functionalized graphene enables highly Wu, B., Cui, J., Fang, X., Xie, Y., and Zheng, N.
64. Tian, L., Luan, J., Liu, K.K., Jiang, Q., Tadepalli, efficient solar thermal steam generation. ACS (2018). A cake making strategy to prepare
S., Gupta, M.K., Naik, R.R., and Singamaneni, Nano 11, 5510–5518. reduced graphene oxide wrapped plant fiber
S. (2016). Plasmonic biofoam: a versatile sponges for high-efficiency solar steam
optically active material. Nano Lett. 16, 78. Awad, F.S., Kiriarachchi, H.D., AbouZeid, generation. J. Mater. Chem. A 6, 14571–
609–616. K.M., Özgür, Ü., and El-Shall, M.S. (2018). 14576.
Plasmonic graphene polyurethane
65. Wang, X., He, Y., Liu, X., Shi, L., and Zhu, J. nanocomposites for efficient solar water 91. Chen, C., Li, Y., Song, J., Yang, Z., Kuang, Y.,
(2017). Investigation of photothermal heating desalination. ACS Appl. Energy Mater. 1, Hitz, E., Jia, C., Gong, A., Jiang, F., Zhu, J.Y.,
enabled by plasmonic nanofluids for direct 976–985. et al. (2017). Highly flexible and efficient solar

716 Joule 3, 683–718, March 20, 2019


steam generation device. Adv. Mater. 29, 104. Wang, Y., Liu, H., Chen, C., Kuang, Y., Song, Zhao, L.-D. (2018). A mimetic transpiration
1701756. J., Xie, H., Jia, C., Kronthal, S., Xu, X., He, S., system for record high conversion efficiency in
et al. (2018). All natural, high efficient solar steam generator under one-sun. Mater.
92. Wang, Y., Zhang, L., and Wang, P. (2016). Self- groundwater extraction via solar steam/vapor Today Energy 8, 166–173.
floating carbon nanotube membrane on generation. Adv. Sustain. Syst. 2018, 1800055.
macroporous silica substrate for highly 117. Ni, G., Li, G., Boriskina, S.V., Li, H., Yang, W.,
efficient solar-driven interfacial water 105. Ma, S., Chiu, C.P., Zhu, Y., Tang, C.Y., Long, Zhang, T., and Chen, G. (2016). Steam
evaporation. ACS Sustain. Chem. Eng. 4, H., Qarony, W., Zhao, X., Zhang, X., Lo, W.H., generation under one sun enabled by a
1223–1230. and Tsang, Y.H. (2017). Recycled waste black floating structure with thermal concentration.
polyurethane sponges for solar vapor Nat. Energy 1, 16126.
93. Wang, X., He, Y., Cheng, G., Shi, L., Liu, X., generation and distillation. Appl. Energy 206,
and Zhu, J. (2016). Direct vapor generation 63–69. 118. Yang, Y., Yang, X., Fu, L., Zou, M., Cao, A., Du,
through localized solar heating via carbon- Y., Yuan, Q., and Yan, C.-H. (2018). Two-
nanotube nanofluid. Energy Convers. 106. Yin, Z., Wang, H., Jian, M., Li, Y., Xia, K., dimensional flexible bilayer Janus membrane
Manage. 130, 176–183. Zhang, M., Wang, C., Wang, Q., Ma, M., for advanced photothermal water
Zheng, Q.S., et al. (2017). Extremely black desalination. ACS Energy Lett. 3, 1165–1171.
94. Li, T., Liu, H., Zhao, X., Chen, G., Dai, J., Pastel, vertically aligned carbon nanotube arrays for
G., Jia, C., Chen, C., Hitz, E., Siddhartha, D., solar steam generation. ACS Appl. Mater. 119. Yu, S., Zhang, Y., Duan, H., Liu, Y., Quan, X.,
et al. (2018). Scalable and highly efficient Interfaces 9, 28596–28603. Tao, P., Shang, W., Wu, J., Song, C., and
mesoporous wood-based solar steam Deng, T. (2015). The impact of surface
107. Chang, C., Yang, C., Liu, Y., Tao, P., Song, C., chemistry on the performance of localized
generation device: localized heat, rapid water
transport. Adv. Funct. Mater. 28, 1707134.
Shang, W., Wu, J., and Deng, T. (2016). solar-driven evaporation system. Sci. Rep. 5,
Efficient solar-thermal energy harvest driven 13600.
95. Chandrashekara, M., and Yadav, A. (2017). by interfacial plasmonic heating-assisted
Experimental study of exfoliated graphite evaporation. ACS Appl. Mater. Interfaces 8, 120. Wegst, U.G., Bai, H., Saiz, E., Tomsia, A.P.,
23412–23418. and Ritchie, R.O. (2015). Bioinspired structural
solar thermal coating on a receiver with a
materials. Nat. Mater. 14, 23–36.
Scheffler dish and latent heat storage for 108. Wang, Z., Ye, Q., Liang, X., Xu, J., Chang, C.,
desalination. Sol. Energy 151, 129–145. Song, C., Shang, W., Wu, J., Tao, P., and 121. Tao, P., Shang, W., Song, C., Shen, Q., Zhang,
Deng, T. (2017). Paper-based membranes on F., Luo, Z., Yi, N., Zhang, D., and Deng, T.
96. Sajadi, S.M., Farokhnia, N., Irajizad, P., silicone floaters for efficient and fast solar- (2015). Bioinspired engineering of thermal
Hasnain, M., and Ghasemi, H. (2016). Flexible driven interfacial evaporation under one sun. materials. Adv. Mater. 27, 428–463.
artificially networked structure for ambient/ J. Mater. Chem. A 5, 16359–16368.
high pressure solar steam generation. 122. Sun, T., Feng, L., Gao, X., and Jiang, L. (2005).
J. Mater. Chem. A 4, 4700–4705. 109. Zhuang, S., Zhou, L., Xu, W., Xu, N., Hu, X., Li, Bioinspired surfaces with special wettability.
X., Lv, G., Zheng, Q., Zhu, S., Wang, Z., et al. Acc. Chem. Res. 38, 644–652.
97. Liu, Y., Chen, J., Guo, D., Cao, M., and Jiang, (2018). Tuning transpiration by interfacial solar
L. (2015). Floatable, self-cleaning, and carbon- absorber-leaf engineering. Adv. Sci. 5, 123. Song, J., Chen, C., Zhu, S., Zhu, M., Dai, J.,
black-based superhydrophobic gauze for the 1700497. Ray, U., Li, Y., Kuang, Y., Li, Y., Quispe, N.,
solar evaporation enhancement at the air- et al. (2018). Processing bulk natural wood into
water interface. ACS Appl. Mater. Interfaces 110. Gao, X., Ren, H., Zhou, J., Du, R., Yin, C., Liu, a high-performance structural material.
7, 13645–13652. R., Peng, H., Tong, L., Liu, Z., and Zhang, J. Nature 554, 224–228.
(2017). Synthesis of hierarchical graphdiyne-
98. Hao, D., Yang, Y., Xu, B., and Cai, Z. (2018). based architecture for efficient solar steam 124. Chen, C., Song, J., Zhu, S., Li, Y., Kuang, Y.,
Efficient solar water vapor generation generation. Chem. Mater. 29, 5777–5781. Wan, J., Kirsch, D., Xu, L., Wang, Y., Gao, T.,
enabled by water-absorbing polypyrrole et al. (2018). Scalable and sustainable
coated cotton fabric with enhanced heat 111. Ren, H., Tang, M., Guan, B., Wang, K., Yang, approach toward highly compressible,
localization. Appl. Therm. Eng. 141, 406–412. J., Wang, F., Wang, M., Shan, J., Chen, Z., anisotropic, lamellar carbon sponge. Chem 4,
Wei, D., et al. (2017). Hierarchical graphene 544–554.
99. Jiang, Q., Gholami Derami, H.G., Ghim, D., foam for efficient omnidirectional solar-
Cao, S., Jun, Y.-S., and Singamaneni, S. (2017). thermal energy conversion. Adv. Mater. 29, 125. Chen, C., Zhang, Y., Li, Y., Dai, J., Song, J.,
Polydopamine-filled bacterial nanocellulose 1702590. Yao, Y., Gong, Y., Kierzewski, I., Xie, J., and
as a biodegradable interfacial photothermal Hu, L. (2017). All-wood, low tortuosity,
evaporator for highly efficient solar steam 112. Ye, M., Jia, J., Wu, Z., Qian, C., Chen, R., aqueous, biodegradable supercapacitors
generation. J. Mater. Chem. A 5, 18397– O’Brien, P.G., Sun, W., Dong, Y., and Ozin, with ultra-high capacitance. Energy Environ.
18402. G.A. (2017). Synthesis of black TiOx Sci. 10, 538–545.
nanoparticles by Mg reduction of TiO2
100. Chen, Q., Pei, Z., Xu, Y., Li, Z., Yang, Y., Wei, Y., nanocrystals and their application for solar 126. Wheeler, T.D., and Stroock, A.D. (2008). The
and Ji, Y. (2018). A durable monolithic water evaporation. Adv. Energy Mater. 7, transpiration of water at negative pressures in
polymer foam for efficient solar steam 1601811. a synthetic tree. Nature 455, 208–212.
generation. Chem. Sci. 9, 623–628. 127. Xu, N., Hu, X., Xu, W., Li, X., Zhou, L., Zhu, S.,
113. Chala, T.F., Wu, C.M., Chou, M.H., and Guo,
Z.L. (2018). Melt electrospun reduced and Zhu, J. (2017). Mushrooms as efficient
101. Jiang, F., Liu, H., Li, Y., Kuang, Y., Xu, X., Chen,
tungsten oxide/polylactic acid fiber solar steam-generation devices. Adv. Mater.
C., Huang, H., Jia, C., Zhao, X., Hitz, E., et al.
membranes as a photothermal material for 29, 1606762.
(2018). Lightweight, mesoporous, and highly
light-driven interfacial water evaporation.
absorptive all-nanofiber aerogel for efficient 128. Wang, X., He, Y., Liu, X., Cheng, G., and Zhu,
solar steam generation. ACS Appl. Mater. ACS Appl. Mater. Interfaces 10, 28955–28962.
J. (2017). Solar steam generation through bio-
Interfaces 10, 1104–1112. 114. Li, Y., Gao, T., Yang, Z., Chen, C., Luo, W., inspired interface heating of broadband-
Song, J., Hitz, E., Jia, C., Zhou, Y., Liu, B., et al. absorbing plasmonic membranes. Appl.
102. Xue, G., Liu, K., Chen, Q., Yang, P., Li, J., Ding, (2017). 3D-printed, all-in-one evaporator for Energy 195, 414–425.
T., Duan, J., Qi, B., and Zhou, J. (2017). Robust high-efficiency solar steam generation under
and low-cost flame-treated wood for high- 1 sun illumination. Adv. Mater. 29, 1700981. 129. Ni, G., Zandavi, S.H., Javid, S.M., Boriskina,
performance solar steam generation. ACS S.V., Cooper, T.A., and Chen, G. (2018). A salt-
Appl. Mater. Interfaces 9, 15052–15057. 115. Li, X., Lin, R., Ni, G., Xu, N., Hu, X., Zhu, B., Lv, rejecting floating solar still for low-cost
G., Li, J., Zhu, S., and Zhu, J. (2018). Three- desalination. Energy Environ. Sci. 11, 1510–
103. Liu, H., Chen, C., Chen, G., Kuang, Y., Zhao, dimensional artificial transpiration for efficient 1519.
X., Song, J., Jia, C., Xu, X., Hitz, E., Xie, H., solar waste-water treatment. Natl. Sci. Rev. 5,
et al. (2018). High-performance solar steam 70–77. 130. Zhao, J., Yang, Y., Yang, C., Tian, Y., Han, Y.,
device with layered channels: artificial tree Liu, J., Yin, X., and Que, W. (2018). A
with a reversed design. Adv. Energy Mater. 8, 116. Liu, P.-F., Miao, L., Deng, Z., Zhou, J., Su, H., hydrophobic surface enabled salt-blocking
1701616. Sun, L., Tanemura, S., Cao, W., Jiang, F., and 2D Ti3C2 MXene membrane for efficient and

Joule 3, 683–718, March 20, 2019 717


stable solar desalination. J. Mater. Chem. A 6, of An Urbanizing World (IntechOpen). https:// oxide leaf for solar desalination with zero
16196–16204. doi.org/10.5772/intechopen.72876. liquid discharge. Environ. Sci. Technol. 51,
11701–11709.
131. Kashyap, V., Al-Bayati, A., Sajadi, S.M., 143. Zhang, P., Li, J., Lv, L., Zhao, Y., and Qu, L.
Irajizad, P., Wang, S.H., and Ghasemi, H. (2017). Vertically aligned graphene sheets 155. Shi, Y., Zhang, C., Li, R., Zhuo, S., Jin, Y., Shi, L.,
(2017). A flexible anti-clogging graphite film membrane for highly efficient solar thermal Hong, S., Chang, J., Ong, C., and Wang, P.
for scalable solar desalination by heat generation of clean water. ACS Nano 11, (2018). Solar evaporator with controlled salt
localization. J. Mater. Chem. A 5, 15227– 5087–5093. precipitation for zero liquid discharge
15234. desalination. Environ. Sci. Technol. 52, 11822–
144. Solargis imaps global horizontal solar
11830.
132. Zhao, J., Cheng, H., Wang, X., Cheng, W., and irradiations. http://www.solargis.info/doc/
Cheng, F. (2018). Experimental investigation _pics/freemaps/1000px/ghi/SolarGIS-Solar-
156. Xue, G., Xu, Y., Ding, T., Li, J., Yin, J., Fei, W.,
and cost assessment of the salt production by map-World-map-en.png.
Cao, Y., Yu, J., Yuan, L., Gong, L., et al. (2017).
solar assisted evaporation of saturated brine.
145. Li, G., Law, W.-C., and Chan, K.C. (2018). Water-evaporation-induced electricity with
Chin. J. Chem. Eng. 26, 701–707.
Floating, highly efficient, and scalable nanostructured carbon materials. Nat.
133. Li, X., Li, J., Lu, J., Xu, N., Chen, C., Min, X., graphene membranes for seawater Nanotechnol. 12, 317–321.
Zhu, B., Li, H., Zhou, L., Zhu, S., et al. (2018). desalination using solar energy. Green Chem.
Enhancement of interfacial solar vapor 20, 3689–3695. 157. Zhu, L., Gao, M., Peh, C.K.N., Wang, X., and
generation by environmental energy. Joule 2, Ho, G.W. (2018). Self-contained monolithic
1331–1338. 146. Zhou, X., Zhao, F., Guo, Y., Zhang, Y., and Yu, carbon sponges for solar-driven interfacial
G. (2018). A hydrogel-based antifouling solar water evaporation distillation and electricity
134. Wang, Y., Wang, C., Song, X., Huang, M., evaporator for highly efficient water generation. Adv. Energy Mater. 8, 1702149.
Megarajan, S.K., Shaukat, S.F., and Jiang, H. desalination. Energy Environ. Sci. 11, 1985–
(2018). Improved light-harvesting and thermal 1992. 158. Gao, F., Li, W., Wang, X., Fang, X., and Ma, M.
management for efficient solar-driven water (2016). A self-sustaining pyroelectric
evaporation using 3D photothermal cones. 147. Yang, Y., Zhao, H., Yin, Z., Zhao, J., Yin, X., Li, nanogenerator driven by water vapor. Nano
J. Mater. Chem. A 6, 9874–9881. N., Yin, D., Li, Y., Lei, B., Du, Y., et al. (2018). A Energy 22, 19–26.
general salt-resistant hydrophilic/
135. Zhang, P., Liao, Q., Yao, H., Cheng, H., Huang, hydrophobic nanoporous double layer design 159. Li, X., Min, X., Li, J., Xu, N., Zhu, P., Zhu, B.,
Y., Yang, C., Jiang, L., and Qu, L. (2018). for efficient and stable solar water Zhu, S., and Zhu, J. (2018). Storage and
Three-dimensional water evaporation on a evaporation distillation. Mater. Horiz. 5, 1143– recycling of interfacial solar steam enthalpy.
macroporous vertically aligned graphene 1150. Joule 2, 2477–2484.
pillar array under one sun. J. Mater. Chem. A
6, 15303–15309. 148. Yang, X., Yang, Y., Fu, L., Zou, M., Li, Z., Cao,
160. Gao, M., Peh, C.K.N., Phan, H.T., Zhu, L., and
A., and Yuan, Q. (2018). An ultrathin flexible
Ho, G.W. (2018). Solar absorber gel: localized
136. Hong, S., Shi, Y., Li, R., Zhang, C., Jin, Y., and 2D membrane based on single-walled
macro-nano heat channeling for efficient
Wang, P. (2018). Nature-inspired, 3D origami nanotube-MoS2 hybrid film for high-
plasmonic Au nanoflowers photothermic
solar steam generator toward near full performance solar steam generation. Adv.
vaporization and triboelectric generation.
utilization of solar energy. ACS Appl. Mater. Funct. Mater. 28, 1704505.
Adv. Energy Mater. 8, 1800711.
Interfaces 10, 28517–28524.
149. Dongare, P.D., Alabastri, A., Pedersen, S.,
137. Song, H., Liu, Y., Liu, Z., Singer, M.H., Li, C., Zodrow, K.R., Hogan, N.J., Neumann, O., Wu, 161. Neumann, O., Neumann, A.D., Tian, S.,
Cheney, A.R., Ji, D., Zhou, L., Zhang, N., Zeng, J., Wang, T., Deshmukh, A., Elimelech, M., Thibodeaux, C., Shubhankar, S., Müller, J.,
X., et al. (2018). Cold vapor generation et al. (2017). Nanophotonics-enabled solar Silva, E., Alabastri, A., Bishnoi, S.W.,
beyond the input solar energy limit. Adv. Sci. membrane distillation for off-grid water Nordlander, P., et al. (2017). Combining solar
5, 1800222. purification. Proc. Natl. Acad. Sci. USA 114, steam processing and solar distillation for fully
6936–6941. off-grid production of cellulosic bioethanol.
138. Ding, H., Peng, G., Mo, S., Ma, D., Sharshir, ACS Energy Lett. 2, 8–13.
S.W., and Yang, N. (2017). Ultra-fast vapor 150. Fujiwara, M., and Kikuchi, M. (2017). Solar
generation by a graphene nano-ratchet: a desalination of seawater using double-dye- 162. Ravaghi-Ardebili, Z., Manenti, F., Corbetta,
theoretical and simulation study. Nanoscale 9, modified PTFE membrane. Water Res. 127, M., Pirola, C., and Ranzi, E. (2015). Biomass
19066–19072. 96–103. gasification using low-temperature solar-
driven steam supply. Renew. Energy 74,
139. Kieu, H.T., Liu, B., Zhang, H., Zhou, K., and 151. Li, S.-F., Liu, Z.-H., Shao, Z.-X., Xiao, H.-s., and 671–680.
Law, A.W.-K. (2018). Molecular dynamics Xia, N. (2018). Performance study on a passive
study of water evaporation enhancement solar seawater desalination system using 163. Alvarez, P.J.J., Chan, C.K., Elimelech, M.,
through a capillary graphene bilayer with multi-effect heat recovery. Appl. Energy 213, Halas, N.J., and Villagrán, D. (2018). Emerging
tunable hydrophilicity. Appl. Surf. Sci. 452, 343–352. opportunities for nanotechnology to enhance
372–380. water security. Nat. Nanotechnol. 13,
152. Xue, G., Chen, Q., Lin, S., Duan, J., Yang, P.,
634–641.
140. Peng, G., Ding, H., Sharshir, S.W., Li, X., Liu, Liu, K., Li, J., and Zhou, J. (2018). Highly
H., Ma, D., Wu, L., Zang, J., Liu, H., Yu, W., efficient water harvesting with optimized solar
164. Zhang, Y., Zhao, D., Yu, F., Yang, C., Lou, J.,
et al. (2018). Low-cost high-efficiency solar thermal membrane distillation device. Global
Liu, Y., Chen, Y., Wang, Z., Tao, P., Shang, W.,
steam generator by combining thin film Chall. 2.
et al. (2017). Floating rGO-based black
evaporation and heat localization: both
153. Chiavazzo, E., Morciano, M., Viglino, F., membranes for solar driven sterilization.
experimental and theoretical study. Appl.
Fasano, M., and Asinari, P. (2017). Passive Nanoscale 9, 19384–19389.
Therm. Eng. 143, 1079–1084.
high-yield seawater desalination at below one
141. UN-Water. (2018). Water scarcity. http://www. sun by modular and low-cost distillation. 165. Li, J., Du, M., Lv, G., Zhou, L., Li, X., Bertoluzzi,
unwater.org/water-facts/scarcity/. https://arxiv.org/abs/1702.05422. L., Liu, C., Zhu, S., and Zhu, J. (2018). Interfacial
solar steam generation enables fast-
142. Neto, S. (2018). Territorial integration of water 154. Finnerty, C., Zhang, L., Sedlak, D.L., Nelson, responsive, energy-efficient, and low-cost off-
management in the city. In Water Challenges K.L., and Mi, B. (2017). Synthetic graphene grid sterilization. Adv. Mater. 30, 1805159.

718 Joule 3, 683–718, March 20, 2019

You might also like