Next Article in Journal
Research on a Framework for Sustainable Campus Eco-Architecture Selection: Taking a Taiwan High School as an Example
Previous Article in Journal
Energy-Efficient Scheduling in Job Shop Manufacturing Systems: A Literature Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Water Harvesting: A Review of Materials, Devices and Applications

1
Division of Natural and Applied Sciences, Duke Kunshan University, Kunshan 215316, China
2
Environmental Research Center, Duke Kunshan University, Kunshan 215316, China
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(10), 6244; https://0-doi-org.brum.beds.ac.uk/10.3390/su14106244
Submission received: 7 April 2022 / Revised: 13 May 2022 / Accepted: 18 May 2022 / Published: 20 May 2022
(This article belongs to the Special Issue Advances in New Technologies of Water Harvesting and Sustainability)

Abstract

:
Water is essential for life. However, water scarcity is becoming one of the most severe issues worldwide in terms of its potential impacts. There are diverse forms of water on earth and water harvesting from them is quite feasible to access more fresh water for drinking, sanitation and irrigation. In this review, we summarize the recent technologies of various water harvesters, based on different forms of water resources, aiming to improve the water harvesting systems. We mainly address three points: forming principles of different water circumstance, working mechanism of typical water harvesters, and the challenges and future research orientations. This systemic review on recent technologies in water harvesting provides insight into the sustainable water resources, water supply, and water collecting systems for the future.

1. Introduction

Water is essential to life on Earth, as an indispensable substance for stabilizing the biological environment within carbon-based life owing to its ideal physical, chemical, and biological properties [1]. Despite of the tremendous amount of water on Earth, only 2.5% is fresh water, mainly stored in glaciers and ice caps. Within this, there is only 1.2% of fresh water stored on the Earth’s surface [2]. The limited fresh water resources and their uneven distribution have led to water crises in many regions, and it is becoming more emergent with the increasing demand of water for industrial, agricultural, and domestic consumption [3,4]. Therefore, how to relieve water shortage has become a hot topic in recent years and is expected to be a serious issue in the next few decades. Goal 6 of the Sustainable Development Goals (SDGs) proposed by United Nations in 2015 also highlights the high demand for safe water accessibility, especially in regions where water shortage is severe. To fundamentally meet the demand of fresh water, attention has been drawn to water resources which are not directly accessible such as water vapor and seawater. Technologies that harvest water from atmosphere or oceans have therefore been developed and proven to be a promising solution to water shortage in the past few decades [5,6,7,8,9]. The various sources or forms of water on earth make harvesting process generally available under different environmental conditions.
Water harvesting technology has been studied since the last century, and great merits have been achieved. The design principles of water harvesting technology are commonly inspired by nature. The back of desert beetles possess alternating hydrophilic bumps and hydrophobic surfaces which enable them to harvest water in dry deserts [10]. The special aligned structures of spines, barbs, and trichomes on the cactus, a plant famous for its vitality in deserts, allow directional transportation of water to their roots [11]. The alternating structures of spindle-knots and joints on spider silk also demonstrate an illuminating idea of directional water transportation [12]. A deep understanding of the mechanism will contribute to the development of bio-inspired materials and devices with higher water harvesting efficiency and compatible integrated water harvesting systems. In addition, the current water harvesting technology all over the world is based on several mature mechanisms such as urban domestic rainwater harvesting systems, reverse osmosis (RO), multi-stage flash (MSF), and multi-effect distillation (MED). Research on improving the water harvesting efficiency of these devices or equipment by integrating with new materials or designs is realistic and significant in support of the development of the whole water harvesting industry.
In this review, we summarize the recent advances in water harvesting technology categorized by the different forms of water on earth, i.e., fog, rain, and dew, aiming to draw global attention to the water crisis and inspire future designs based on the local conditions. In each subsection, the formation and distribution are introduced, demonstrating the environmental conditions necessary for certain types of water harvesting. Different designs of water harvesters and newly developed materials for the specific water harvesting category are discussed. Finally, the recent technology of water harvesting is commented on and analyzed based on the limitations and restrictions of its practical application, and the future perspectives of each technology category are also suggested. This work will provide a meaningful insight into the new design of water harvesters with a higher collection efficiency, to relieve the water crisis in our world.

2. Water Harvesting Technologies

2.1. Water Harvesting from Fog

Fog is a common natural phenomenon defined as a cloud made up of abundant water droplets close to the ground, the diameters of which typically range from 1–30 μ m, reducing the visibility below one kilometer [13,14]. Fog can form in multiple processes, such as advection of moist over cold water bodies and the adiabatic cooling of moist water during uphill transport [15]. Since oceans are filled with tremendous water content, fog can easily form near the coastline especially combined with mountainous areas. Figure 1 marked these costal locations with plenty of foggy days to implement the fog harvesting technology. Even though fog water might contain marine salt or dust, it is expected to meet the drinking water standards after simple treatment [14,16]. Therefore, water harvesting from fog has been proposed as a feasible solution to relieving water shortage, especially in areas where fog is prevalent but rainfall is scarce, such as the fog harvesting projects achieved in Chile [17].
The research on fog harvesting has been developed from traditional mesh designs to state-of-art collectors with bio-inspired materials to achieve the advanced functions and high performance. Initial fog harvesters were designed as simply as a single or double layers of vertical meshes equipped with a water tank, given the horizontal traveling pattern of fog in the atmosphere [5,18]. Materials such as polypropylene were used in practical applications at an early time. The standard fog collector (SFC) (Figure 2a) was proposed, further developed and fabricated based on the specific environmental conditions, in order to relieve local water shortage [19,20,21]. For example, Schemenauer [19] reported one of the earliest SFCs with an average water collection rate of 1.26 L m 2 h 1 located in El Tofo, Chile. In the past two decades, the natural biological structures with advanced capability of fog capture, coalescence, and directional transportation, attracted researchers’ attention to spark new ideas and designs. For example, the coexistence of hydrophobic and hydrophilic surfaces on the back enables desert beetles to harvest fog water from the air [10]. Inspired by the alternating pattern, Wang et al. [22] reported a superhydrophobic-superhydrophilic hybrid (SSH) fabric coated by Co, Fe and n-octadecyl thiol (Figure 2b). The fabric with a Co and Fe concentration ratio of 3:1 was tested to achieve the fog collection rate of 0.52 L m 2 h 1 under the laboratory conditions, while the water collection efficiency was only 0.16%. The special aligned structures of spines, barbs and trichomes on cactus sparked a new idea of directional water transportation [11]. Cao et al. [23] fabricated a cactus-inspired fog collector with the approximate diameter of 2 cm, by integrating cactus spine-like hydrophobic conical micro-tip arrays with the hydrophilic cotton matrix (Figure 2c). At a normal fog velocity of 45–50 cm/s, one such cactus-inspired fog collector was able to harvest about 3 mL water per 10 min. Additionally, some advanced harvesters were inspired by spider silk with the special alternating structures of spindle-knots and joints (Figure 2d), to achieve the directional water transportation [12]. The artificial spider silk was fabricated to mimic the structure of wet rebuilt spider silk and proved to show its directional water collection capability.
Recent research has developed more advanced bio-inspired materials in order to resolve problems in practical applications. A slippery liquid-infused porous surface (SLIPS) proposed by Wong et al. [24] in 2011, achieved omniphobicity, high-pressure stability, and self-healing capability, compared with a traditional superhydrophobic surface (SHS). SLIPS is beneficial to water coalescence and transportation, raising the interest of groups of researchers. The latest research proposed several polished structures to improve the applicability of SLIPS. Feng et al. [25] recently infused SHS with silicone oil to create a SLIPS with hierarchical micro-/nanostructures (MNS-SLIPS) shown in Figure 3a. MNS-SLIPs successfully reduced lubricant loss and improved the water collection rate, up to 8.52 L m 2 h 1 , more efficient than the traditional design of SLIPS. Han and Guo [26] fabricated a three-dimensional ball-cluster structure used as the substrate of lubricate-infused surface (Figure 3b). Due to the inner 3D grid system and the supportive nano-structures, the new design was endowed with a good durability and achieved a water collection rate of approximately 45.18 L m 2 h 1 .
Previous research aiming at improving fog harvesting efficiency was mainly focused on the development of one- or two-dimensional materials. However, the 1D or 2D structures are not ideal enough within the real fog environment. Although fewer investigations have been conducted on 3D fog harvesting designs, the interest in this specific direction is rising. Early in 2011, Andrews et al. [27] reported the 3D hierarchical structures of a plant, Cotula fallax (Figure 4a), and demonstrated that the compact structures greatly assisted in water harvesting from fog, where such 3D structures increased the surface-to-volume ratio and hence potentially improved the efficiency of water capture. The air plasma treatment changed the surface chemistry of the plant and further increased the water collection efficiency. Lee et al. [28] recently reported a new cactus-inspired 3D fog harvester (Figure 4b) integrated with a developed temperature-responsive material, poly-(N-isopropylacrylamide) (PNIPAAm). The PNIPAAm enabled the interchange of absorbing and releasing water via the interpenetrating polymer network (IPN) hydrogel, in response to the temperature change which has a potential application value in water management. In addition, the omnidirectional water capture of the design contributed to improving water capture rate, up to 2.09 L m 2 h 1 , higher than that of the superhydrophilic IPN hydrogel, superhydrophobic copper mesh (SHPM), and pristine copper mesh (PTM).
The integration of multiple bio-inspired features in the design and fabrication of materials or devices for fog harvesting has been hot topics. Inspired by Namib desert beetles and leaf veins, Yu et al. [29] reported a dual-biomimetic knitted fabric (SHL-SHB@TF) fabricated by modern weaving technology (Figure 5a). SHL–SHB@TF is featured for its excellent stretching, anti-aging, anti-bacterial, and water capture ability. Wang et al. [30] successfully fabricated an integrated bio-inspired surface (IBS) by combining features of beetles, spider silk, cactus, and Sarracenia (Figure 5b), by an innovative capillary-induced selective oxidation method. To engage SLIPS in a complete fog harvesting process, Feng et al. [31] developed a quadruple-biomimetic surface (SLIPS-SHIS) inspired by desert beetles, pitcher plants, cactus and Nephrolepis auriculata (Figure 5c). SLIPS-SHIS contained all four fog-harvest processes, including fog capture, coalescence, transportation, and collection; together, the synergetic functions of the upgraded processes improved the overall fog harvesting efficiency.
Although bio-inspired materials raised much attention regarding the subject of improving the fog-harvesting rate, inventions regarding the physical structure of fog harvesters remained a promising solution to difficulties induced by the application of new materials. Given the blockage issue of traditional net-like meshes, Shi et al. [32] invented fog harps, comprised of vertical fibers (Figure 6a), where the acrylic frames were cut by laser and the water drainage was fabricated by 3D-printing. In the lab testing, the fog harps with the stainless steel wires could harvest up to 3 times more water compared to the traditional mesh netting of the same size. Considering the similar shortcomings, Cheng et al. [33] proposed an arc-shape fog catcher brushes (Figure 6b) made from aluminum foil and coated with polydimethylhydrosiloxane (PMHS) to achieve a better performance of water collection. The arc-shape brushes with the unit pattern width of 3.2 mm had the best fog-harvesting rate of 4.45 L m 2 h 1 , in their study.
The development of fog harvesting techniques was rapid in the past few years. More advanced and feasible designs of new materials, novel structures, and innovative devices have been continuously investigated in order to improve collection efficiency. However, imperfections exist in fog harvesting technologies. Firstly, most of the work done by now were carried out in laboratory conditions while the real fog environment can be more complicated. Ultraviolet radiations from sunlight, temperature changes, air pollutants, extreme weather and other environmental factors are likely to affect the actual efficiency of current designs. Secondly, even though some novel designs showed a greater water capture efficiency, limitations of certain properties may become significant issues in practical applications. For example, the cactus-inspired 3D fog harvester reported by Lee et al. [28] can work more than ten times in response to temperature change before it suffers from morphological deformation. Nevertheless, their durability outdoors may need investigating, given the cost of maintenance. Thirdly, apart from improving the quality of a single part of the fog harvesting process, investigations on the integration of components with different functions are necessary to achieve overall high efficiency. For example, a well-designed component for water capture may not fit into a device for efficient water transportation as a whole piece of water harvester. Fourthly, the quality of water from fog harvesting should be considered. Since fresh water is of high demand, the quality of harvested water is expected to have a criterion as important as efficiency when evaluating a new device in the future. Last but not least, beyond the current scope as it might be, the mass production of a prototype also needs more attention, and the possible points include economic manufacture technology and maintenance, the cost of raw materials and labor. Overall, fog harvesting technology is a promising solution to relieving water shortage in the regions where fog is plentiful but rain is rare, indispensable in the family of water harvesters.

2.2. Water Harvesting from Dew

Dew is defined as water droplets condensed from vapor in the air onto a surface whose temperature is lower than the dew point [13]. The formation of dew normally goes through the two main steps: nucleation and growth [34]. Heterogeneous nucleation (Figure 7b), defined as condensation of water vapor on a substrate, can lower or even suppress the energy barriers compared with homogeneous nucleation (Figure 7a) [34]. Therefore, heterogeneous nucleation is more commonly observed than homogeneous nucleation in life. Given the widespread formation of dew [35], more research on water harvesting from dew has been one of hot topics in the past few years, as an option to resolve global water shortage.
The traditional design of dew harvesting can date back to more than two thousands years ago when the artificial condensers—aerial wells were built [37], attributed to the temperature difference between the voids on the stone walls and the incoming air [38]. With regard to the formation conditions required for dew, rapid nucleation at a surface is the first and the essential step to achieve efficient dew harvesting. The first key factor influencing the nucleation rate is the temperature difference between the substrate ( T s ) and the surrounding atmosphere ( T r ) [34]. Due to this characteristic of dew formation, the design idea of efficiently cooling down the surface below the dew point is proposed. Passive radiative cooling could be the most straightforward methodology that does not need high-technology or external energy input [38]. The rationale of the technology is that the emitting infrared radiation from a surface at night tends to exceed the radiation absorption from the ambient environment, resulting in the lower surface temperature than that of the surroundings [39]. When daytime comes and more water moisture enters the atmosphere, dew forms at the surface, with this idea motivating research on low emissivity materials [6]. Alnaser and Barakat [40] tested the dew collection efficiency of aluminum, glass and polyethylene foils, respectively, and reported that aluminum achieved the highest collection rate of 1.3 L m 2 h 1 among the three types of materials. As more interest was cast on the passive radiative cooling, different types of designs were quickly developed, i.e., film-based, nanoparticle-based, or photonic radiators, contributing to a more efficient cooling process [41]. Given the low yield and their reliance on environmental conditions for passive radiative cooling, active condensation also attracted people’s attention, although external energy is normally input in the systems [42]. The traditional design of active condensation is to keep the surface cool for a longer time than passive ways by integrating the surface with a cooling system. Apart from passive radiative cooling and active condensation, designs combining dew condensation with water extraction from the air were integrated to achieve a higher water harvesting efficiency [6].
Surface property is another key factor that would impact the dew harvesting efficiency. Hydrophilic surfaces were proposed and designed as the condensation substrates in early studies. By testing the dew condensation ability on surfaces with the alternating hydrophobic and hydrophilic regions, Varanasi et al. [43] found that water droplets would preferentially condense on the hydrophilic regions due to the lower energy barriers (Figure 8). Lee et al. [44] also reported that hydrophilic surfaces could condense approximately 30 mg more water per hour compared with moderately hydrophobic surfaces, whose surface areas were both 9 cm 2 . The observation was consistent with the dew harvesting behaviour of green tree frogs, Litoria caerulea, living in northern Australia [45]. Green tree frogs are living without the direct access to the nearby water resources because the hydrophilic skin cooled at night can nucleate water droplets when frogs come back to the warm humid tree hollows [44]. Instead of drinking water, frogs can directly absorb condensed water by skin [46]. Despite the remarkable capture ability at hydrophilic surfaces, they were still not the ideal design in the application of dew harvesting due to the low mobility of water droplets on such surfaces. Rapid removal of water droplets on the surface hence became the crucial step for continuous dropwise condensation. To solve this problem, superhydrophobic surfaces turned out to be a promising candidate [47]. Inspired by lotus leaves, Chen et al. [47] designed a two-tier surface which can remain superhydrophobic before and after dew condensation. The structure enabled water droplets to stand in a sustained Cassie’s state and have a higher mobility. As the nanotechnology developed, opportunities were provided for researchers to fabricate more micro/nanostructures to achieve the superhydrophobic property, which raised great interest in the past years [48,49]. However, even though the advent of superhydrophobic surfaces increased the overall dew collection rate via rapid water transportation compared with hydrophilic surfaces, the (super)hydrophobic surfaces generally generate a high energy barrier for nucleation which is unfavourable for capturing water vapour [50]. Aware of the dilemma, Hou et al. [50] developed biphilic structured surfaces composed of nanoscale hydrophilic bumps and superhydrophobic substrates inspired by desert beetles and lotus effect (Figure 9a). When the water droplets were condensed and coalesced on the hydrophilic bumps, the droplets would be ejected by the superhydrophobic surface in time. The biphilic surface was reported to have 349% water collection rate and 184% heat transfer coefficient approximately as compared to the state-of-the-art superhydrophobic surface in a dry environment. In a recent research, Hou et al. [51] designed a superhydrophilic/superhydrohobic hybrid surfaces by using laser ablation (Figure 9b). With a spacing of 1.5 mm between triangles, the water collection rate of the hybrid surfaces was measured to be 1.107 ± 0.057 L m 2 h 1 , which was about 54% and 21% higher than that of superhydrophobic and superhydrophilic surfaces respectively, indicating the great potential of biphilic structured surfaces in increasing the dew harvesting efficiency.
Dew harvesting has the advantage over fog harvesting in that dew requires less atmospheric formation conditions. However, some issues in fog harvesting technologies discussed in the previous section happen to dew harvesting. In addition, there are some distinct drawbacks that make it even farther from practical application. Nioras et al. [35] carried out both theoretical and experimental analysis on the water collection rate and efficiency summarized in Table 1. Although the efficiency of dew harvesting is approximately two times as large as that of fog harvesting, the absolute water collection rate of dew harvesting is over one order of magnitude smaller. Increasing water collection rate of dew harvesters is hence the essential step currently. As reported by Hou et al. [50], the combination of different wettability by using micro/nanotechnology has proved to be a feasible way to resolve this issue. Apart from improving the efficiency of a separate part, the integrated systems such as combining passive radiative cooling, active cooling or dew harvesting materials of high efficiency together may work as a promising direction and lead to unexpected positive synergetic effects.

2.3. Water Harvesting from Rain

Rain is probably the most well-known type of precipitation in our daily life. In meteorology, rain is defined as liquid water droplets with diameters ranging from 0.5 to 5 mm that fall from the atmosphere [13]. In a global scale, the amount of precipitation varies from place to place. As shown in Figure 10a, Central Africa, South America, Europe, East Asia and Eastern North America are rich of annual precipitation on average while North Africa, Central Asia, Western Asia and Central Australia experience the least rainfall throughout the year. Compared with the global map of water scarcity risks estimated by Gassert et al. [3] (Figure 10b), the overlapping areas of large precipitation and relatively high water scarcity risks such as East Asia indicates the great potential application of rainwater harvesting (RWH) in order to relieve local water shortage. In general, RWH technology can also be applied to other areas where water scarcity risks are high such as Africa even though the low annual precipitation may result in a relatively low yearly yield of collected rainwater.
Rainwater harvesting has a long history for domestic and agricultural purposes, but its unified definition has always been under discussion by the scientific community [53]. Given the widespread classical structure of an RWH system, it is commonly accepted that rainwater harvesting is a process of collecting and storing rainwater for future use, e.g., irrigation, cooking, washing and drinking [8]. The domestic rainwater harvesting (DRWH) in urban and rural areas is one of the most popular RWH technology all over the world. The most common DRWH structure includes the rooftops while other types of catchment are also available [54]. Abdulla and Al-Shareef [55] introduced the importance of rooftops for RWH set-up in Jordan, one of the most water-stressed semi-arid countries in the world. The typical RWH system used in Jordan was composed of a collection roof, a conveyance system and a storage tank (Figure 11a). It was reported that a maximum of 15.5 Mm 3 every year of rainwater could be collected if all rooftop surfaces were used, which was equivalent to 5.6% of the total domestic water supply in 2005. The water quality of rooftop-collected water was also tested and was considered to be drinkable in a usual case. Melville-Shreeve et al. [56] designed a similar set-up of rooftop rainwater harvesters used in UK. Based on the fundamental structures, more complicated innovative RWH systems were integrated in practical applications for a more intelligent control of rainwater supply. Figure 11b demonstrated a real-time control system used in Australia, Korea and USA that enables the control of the storage volume of water in the tank based on the weather forecast data so that enough capacity can be available if a storm hits.
Rainwater harvesting technologies are also well adopted in agriculture such as irrigations. The proper design of an RWH system can greatly improve the rainwater-use efficiency and sustain the rainfed crops [57]. Xiao et al. [58] compared two types of micro-catchment rainwater harvesting systems combined with supplemental irrigation used in a semiarid region of China. According to the experimental data, the field cultivation of rainwater harvesting with a sowing in the furrow between film-covered ridges (SFFCR) (Figure 12a) and with a sowing in the holes on film-covered ridges (SHFCR) (Figure 12b) can save water at the efficiency of 5.5–5.8% and 9.4–9.6% respectively, which showed that the combination of water harvesting and supplemental irrigation can play a significant role in improving the crop yield and water use in this region. Apart from directly increasing the rainwater capture efficiency by soil and plants, storage of rainwater is also common in agriculture. Bruins et al. [59] introduced an easy design of the liman terrace used in Kenya, a pond that can store runoff originated from rain (Figure 12c). Interestingly, limans can also be built in a series and water will fill out limans one by one by transporting through the spillway when runoff is of a large amount. Aware of the shortcomings in traditional ponds, Pari et al. [60] recently proposed an innovative flexible water storage system (FWSS) for macro-catchment (Figure 12d). FWSS can be easily folded and moved, and no concrete base was needed for installation. FWSS turned out to have lower costs (€16.94/m 3 vs. €20.41/m 3 ) and lower environmental impacts (17.04 g per m 3 CO 2 vs. 28.2 g per m 3 CO 2 ) than traditional ponds, for which FWSS can be a good alternative component of the rainwater harvesting system in agriculture.
The recent research on rainwater harvesting starts to develop a more integrated method and technology in order to further improve the quality of water and the efficiency of enlarging the scale of rainwater harvesting. Given that most research on RWH was focused on non-potable usage, Alim et al. [61] developed an integrated rainwater harvesting unit (IRWHU) (Figure 13a). IRWHU adopted classical rooftop rainwater harvesting system while combined with a filtration unit that could generate drinkable water, capable of producing drinking water at a capacity of 348 ± 20 L/day. In China, Sponge City Concept (SCC) is proposed to mimic the natural hydrological cycle to absorb, infiltrate, store, purify and manage the rainwater. It was planned by the State Council that 20% of urban cities in China should reuse 80% of the rainwater by 2030. Zevenbergen et al. [62] pointed out that SCC is operated widely all over the conutry (Figure 13b). High technology is also integrated into the harvesting systems for a more accurate control of rainwater collection. Adham et al. [63] developed a GIS-based approach to identifying potential sites for rainwater harvesting in Iraq by considering five biophysical factors. The suitability map of potential rainwater harvesting sites (Figure 13c) provided a valuable guideline to optimize the efficiency of RWH technology.
Compared with fog and dew harvesting, the concept of rainwater harvesters is more intuitive given the way of collecting and storage. The simpler design of rainwater harvesting renders this topic more popular in a local scale, and rainfall will primarily determine the yearly yields on average, which makes the comparison of different designs difficult in a global scale. However, even though rainfall and atmospheric conditions vary from place to place, improving the efficiency of rainwater harvesting by learning from the experience from all projects is still valuable while it is important to stay active to modify the design based on the local ecological environment. In addition, the practical applications of rainwater harvesting technology in more places is effective in increasing the overall yield of collected rainwater globally. One key point to achieve this goal is to lower the cost, including materials and labour. Last but not least, collected rainwater still needs purification process for drinking and cooking. With the higher quality, the water may be of greater value in more application fields.

2.4. Water Desalination Technology

Earth is usually called a blue planet because approximately 71% of the surface is covered by ocean. However, only a small proportion of the water on Earth can be directly used by human beings. Shiklamonov [2] reported that only 2.5% of total global water is fresh water, and the rest is not directly accessed mainly due to the high salinity (Figure 14). Given the rapid consumption of fresh water and that fresh water becomes increasingly brackish, desalination, a process of removing salts and other particles from seawater or brackish water, attracts people’s attention as a feasible technology to alleviate water shortage [4]. The major desalination market is along the coastal areas, such as the Gulf Region and the Mediterranean Region, and has been extended to America and Asia [64]. Desalination technology is one of the most well-developed topics in water harvesting during the past decades. The dominant desalination technology includes thermal and membrane-based desalination while many other methods such as chemical-based or absorption desalination are also under investigation [7].
Thermal desalination typically requires input of thermal energy to heat or evaporate seawater for the purpose of separating water and salts. One of the most popular thermal desalination technology is multi-stage flash (MSF) [66] (Figure 15a). Seawater input into the system is heated to 90–110 °C and travels through different stages in which water evaporates from seawater, condenses on the pipe of preheated flowing seawater and is collected as fresh water. There are normally 4 to 40 stages in an MSF unit and 10,000–35,000 m3 of fresh water can be produced per day by one unit. Another traditional thermal desalination technology is multi-effect distillation (MED) [66] (Figure 15b), typically with 2–16 stages. Seawater input is sprayed on horizontal tubes filled with steam, then evaporates and moves into the horizontal tubes of the next stage, collected as fresh water. Meanwhile, the hot water vapor in the horizontal tubes also evaporates seawater spray in this stage, and the whole system is driven by decreasing pressure in different stages. Depending on the length of series, one MED unit could produce fresh water at the capacity of 600–30,000 m3/day.
Given the high energy consumption of the traditional thermal desalination, the renewable energy has raised people’s attention [67]. Solar-thermal desalination is one of clean energy applications, and can be either in direct or indirect way. Indirect solar-thermal desalination is converting solar energy into thermal or electrical energy for driving the traditional thermal desalination systems such as MSF and MED. Direct solar-thermal desalination absorbs the energy from sun for distillation but more suitable for small production. Recently, Lei et al. [68] reported an MXene-decorated 3D honeycomb-fabric design for direct solar-thermal desalination (Figure 15c). Bulk seawater was drawn up through 1D hydrophilic fibrous rods to the patterned surface, and the water can evaporate efficiently by solar energy due to the high light absorbency of MXene and light-trapping ability of honeycomb-fabric structures. The MXene-decorated 3D honeycomb-fabric design demonstrated a high solar efficiency of 93.5% and the evaporation rate of 1.62 kg m−2 h−1. Membrane distillation (MD) (Figure 15d) is also a promising thermally-driven separation method combined with the renewable energy [69,70]. MD is driven by thermally-induced pressure difference between the two sides of the hydrophobic microfiltration membrane which only allows gas (e.g., water vapor) to permeate. Although MD may consume more than ten times of thermal energy than MSF or MED, the low-grade operational temperature made it more compatible to the renewable energy. Other thermal desalination technologies such as compression distillation (MVC), humidification-dehumidification desalination (HDH), solar distillation (SD) have also been gradually developed, for a better performance of thermal desalination [7].
Membrane-based desalination is the dominant desalination technology all over the world (Figure 16), and the main subtypes include reverse osmosis (RO), forward osmosis (FO) and electrodialysis (ED) [71]. RO is the most widespread desalination technology and shares more than half of the total capital of the desalination market as shown in Figure 16 [71]. RO produces fresh water by motivating permeation of pure water through semi-permeable membrane under pressure difference, and the pressure on the feed side commonly ranges from 55–68 bar [64] (Figure 17a). Compared with other desalination technology, RO has the advantage of saving energy and is typically considered to be the most energy-efficient desalination technology so far [72]. ED is a common desalination technology driven by electrical energy, typically composed of anion and cation exchange membranes between which is a spacer gasket [73]. Two electrodes are semi-separated by the membranes. When voltage is applied on the electrodes, reduction and oxidation will occur on the cathode and anode respectively. When seawater enters the ED stack, cations such as Na + will move towards the cathode and anions such as Cl will move towards anode through membranes, leaving fresh water in the stack. The microbial desalination cell (MDC) (Figure 17b) is a novel ED design utilizing microorganisms to convert organic compounds into electricity [74], considered as an eco-friendly and sustainable technology by producing and using bioenergy. FO also evolves membrane-based desalination technology, which draws water from low salinity to high salinity through semi-permeable membranes and involves two main steps: the osmotic dilution of draw solution and extraction of fresh water from diluted draw solution [75]. FO is usually combined with other desalination technology to achieve a more energy-efficient performance. Yuan et al. [76] reported a hybrid system of MDC and FO (Figure 17c) in 2015, showing that MDC-FO had a better performance in removing organic concentration and desalination than standalone MDC, resulting in a lower wastewater volume by 64%.
Apart from thermal and membrane-based desalination technologies which are dominant in the current market, research on other technologies is also on the way. Recently, Guo et al. [77] reported the [emim][Tf2N] ionic liquid, better than the state-of-the-art directional solvent (decanoic acid) in directional solvent extraction (DSE), a developing non-membrane desalination technology (Figure 18a). DSE uses the directional solvent, which can not dissolve in water but can absorb water and reject salts, to separate water from seawater in response to different conditions such as temperature. It was demonstrated that [emim][Tf2N] can yield more than ten times of water than the current decanoic acid. Zhang et al. [78] compared the efficiency of two LiCl supported composite sorbents in adsorption desalination (AD) technology. AD is typically composed of an evaporator, a reactor and a condenser (Figure 18b). Seawater evaporates in the evaporator by heating and temporarily is absorbed by dry sorbents in the reactor. By using solar energy or waste heat, water can be released from the sorbents, enter the condenser and condense as fresh water. AD attracts people’s attention because of several advantages such as its lower corrosion and fouling on tube materials. The predicted specific daily water production (SDWP) of LiCl@SG 30 and LiCl@EVM 45 showed that AD can be a promising desalination technology, and the SDWP of LiCl@SG 30 ranges from 43 to 60 m 3 /(tonne day).
Desalination has been systematically studied in the past few decades, and many technologies have emerged, such as RO, MSF and MED, widely used in practical applications. Improving the performance and reducing the energy required for these technologies may deserve more investigation, but development of combined solar energy collection systems, integrated to current desalination technologies, can play a more significant role. In addition, more types of renewable energy such as wind can also be alternative energy sources, taken into considerations for future designs. Last but not least, the modification on the mature desalination technologies for the application of a wider scale such as desalination of continental brackish water deserves more research.

2.5. Extracting Water from Atmosphere

Approximate 12,900 km 3 of water is stored in the atmosphere [2], especially in the form of water vapor. Aware of the increasing demand of fresh water, scientists started to conceive simple and practical methods to extract water directly from air [79], even in the desert. The idea attracted much attention and sorbent-based water harvesting technology became the popular topic for atmospheric water extraction [80]. The development of sorbents with high water sorption ability, low regeneration cost and fine hydrolytic stability is one of the main directions of research on atmospheric water extraction.
Generally, sorbents can be categorized into absorbents and adsorbents [80]. Absorbents are substances that take up water into their bulk phase [80]. Elashmawy [9] designed a tubular solar still (TSS) using CaCl 2 as the absorbent (Figure 19a), where TSS absorbed water at night and went through regeneration process via solar energy at daytime during which evaporation of water from the absorbent happened. Five separate stills were tested under different wind speeds, and the results showed that the module at 4 m/s air speed had the maximum evaporation and efficiency, 467 mL/m 2 per day and 25%, respectively. Adsorbents are substances with porous structures whose inner surfaces can bind water molecules [80]. Essa et al. [81] designed a double-slope half-cylindrical basin solar still (DS-HCBSS) using silica gel, one of the most popular adsorbents (Figure 19b). In this design, the silica gel was spread on the still with a thickness of 1.5 cm and a parabolic trough solar collector was used to enhance the distillation of water in the morning. As a result, at most 400 mL/m 2 of water per day, can be produced by the DS-HCBSS with silica gel, longitudinal fins, and gravels.
Metal organic frameworks (MOFs) are a unique class of adsorbent, which have the great potential applications in water extraction [80]. MOFs are composed of inorganic cores linked by organic linkers, and the porous structures allow uptake of water molecules. The development of optimal chemical and physical structures of MOFs for water harvesting has become an emerging research topic. Furukawa et al. [82] systematically compared the water adsorption properties of zeolite 13X, BPL carbon, MCM-41 and 20 types of MOFs. The results revealed that MOF-801 and MOF-841 (Figure 20a,b), two Zirconium MOFs, achieved the highest water adsorption performance. Specifically, MOF-801-P realized a water uptake capacity of 280 cm 3 /g at the relative humidity P / P 0 = 0.1 and MOF-841 could take up water at the capacity of 550 cm 3 /g when the relative humidity reached P / P 0 = 0.3 . Due to the excellent water-extraction ability from air, water harvesters based on MOFs materials were further invented. Fathieh et al. [83] fabricated a MOF-based water harvester which went through one adsorption/desorption cycle per day (Figure 20c). The instrument shared the similar idea with TSS and DS-HCBSS while MOFs would not corrode apparatus due to deliquescence such as hygroscopic salts and possess good hydrothermal stability. To enhance the thermalphysical and adsorptive properties of MOF-801, 33 w t % of porous graphite was blended together with MOF-801, and the results demonstrated that 100 g/day of water can be produced by 1 kg MOF-801 in the desert of Arizona USA, illustrating the value of MOF-based water harvester in practical applications. In addition, a newly designed aluminum-based MOF-303 showed around 64.5% higher water production capability in the laboratory conditions than MOF-801. Recently, Hanikel [84] reported a multi-cycle MOF-based water harvester by using photovoltaic module to heat the MOF mildly, achieving multiple adsorption/desorption cycles during day and night (Figure 20d). MOF-303 and three other hydrophilic materials were tested, where MOF-303 even showed better water extracting performance than others, with 1 kg MOF-303 could produce 0.7 L/day of water in Mojave Desert, more efficient than the reported one-cycle MOF-based water harvester.
Water extraction from atmosphere is an ideal technology to obtain fresh water, for air is ubiquitous everywhere. If the technology is developed well, it is striking to produce water in the deserts. However, there exist some issues in current water extraction systems [80]. One key challenge is that the water productivity is still too low to meet the basic demand of water consumption for both active and passive routes. In addition, the current designs of sorbent-based water harvesters are in small scales, far from industrial mass production and practical applications in life. Moreover, the durability of sorbents needs to be largely improved for a sustainable purpose. Although many challenges remain in water extraction technology, especially the applications of MOFs in this scenario, the high performance of new types of adsorbents indicates a bright future of achieving direct water extraction from air in real life.

3. Conclusions

Water is an indispensable resource, and water shortage has brought great threat to life on earth. Water harvesting is a promising solution to the global water crisis, to obtain fresh water from various sources. The concepts of water harvesting technology are commonly inspired by nature such as desert beetles, cactus and spider silk, providing insights into the new design for water collection, transportation and storage. Some mature water harvesting technologies have been widely used in real-life application while research is still in progress in order to develop more options applicable to various environmental conditions and further improve the efficiency.
Water harvesting technology can be categorized into fog harvesting, dew harvesting, rainwater harvesting, desalination and water-extraction from air, based on various forms of water. The corresponding water harvesters have different characteristics: (1) fog droplets suspending in the air can be captured by traveling through net-like meshes or across the bio-inspired materials with special wettability patterns; (2) dew can condense on surfaces whose temperature is lower than the dew point either by active or passive routes. The microstructures of the surface will affect the condensation rate; (3) the design of rainwater harvesters is integrated with storage containers and highly dependent on the local environment. The water harvested is widely used for domestic or agricultural purpose; (4) desalination is probably the most mature water harvesting technology so far for industry-level production. Although various desalination technologies have been developed, RO and MSF are still two main methods in the world; (5) extracting water directly from air is an appealing concept since the expected scope of application is almost everywhere and at any time. MOFs have demonstrated excellent properties of efficient water adsorption and hydrothermal stability, which are promising materials for sorbent-based water harvesters.
Although the state-of-art water harvesting technologies have been achieved with high efficiency, there are still challenges demanding feasible solutions which indicate the future trend of water collection. First of all, the mass production of water collectors is one of the main challenges especially for fog and dew harvesting technologies. Most of water harvesters achieved the high-collection efficiency with only the small prototypes in lab testing. The collection efficiency of the large-scale water harvesters needs further investigations, even in the realistic fog conditions. With the proof of such data, mass production could become possible with cost-effective manufacturing technologies or machinery equipment.
Besides, currently with attention paid on improving the collection efficiency, the quality of water should also be investigated, e.g., drinking, irrigation or sanitation. Reliable water-treatment components integrated to the harvester systems are indispensable to improving the sanitary conditions of fresh water. Additionally, the durability and maintenance of water harvesters should also be taken into account as critical evaluations. This would save the cost for the whole project, and the less maintenance required would popularize the technologies not only in developing countries, but also rural regions. It takes time to overcome these challenges, however, the achievements so far have proven a bright future of water harvesting in alleviating water shortage in a global scale. New designs of water harvesters would continuously arise, and hopefully they will gradually fulfill the gap between the decreasing fresh water resources and the higher demand of water.

Author Contributions

Conceptualization, W.S.; methodology, W.S.; writing–original draft preparation, K.C., Y.T. and W.S.; writing–review and editing, W.S. and K.C.; supervision, W.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. We acknowledge the startup funding support from Duke Kunshan University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Westall, F.; Brack, A. The Importance of Water for Life. Space Sci. Rev. 2018, 214, 1–23. [Google Scholar] [CrossRef]
  2. Shiklomanov, L.A. Water in Crisis: A Guide to the World’s Freshwater Resources; Oxford University Press: Oxford, UK, 1993. [Google Scholar]
  3. Gassert, F.; Luck, M.; Landis, M.; Reig, P.; Shiao, T. Aqueduct Global Maps 2.1: Constructing Decision-Relevant Global Water Risk Indicators; World Resources Institute: Washington, DC, USA, 2014. [Google Scholar]
  4. Service, R.F. Desalination freshens up. Science 2006, 313, 1088–1090. [Google Scholar] [CrossRef] [PubMed]
  5. Abdul-Wahab, S.A.; Lea, V. Reviewing fog water collection worldwide and in Oman. Int. J. Environ. Stud. 2008, 65, 487–500. [Google Scholar] [CrossRef]
  6. Jarimi, H.; Powell, R.; Riffat, S. Review of sustainable methods foratmospheric water harvesting. Int. J. Low-Carbon Technol. 2020, 15, 253–276. [Google Scholar] [CrossRef]
  7. Youssef, P.G.; Al-Dadah, R.K.; Mahmoud, S.M. Comparative analysis of desalination technologies. Energy Preced. 2014, 61, 2604–2607. [Google Scholar] [CrossRef] [Green Version]
  8. Helmreich, B.; Horn, H. Opportunities in rainwater harvesting. Desalination 2009, 248, 118–124. [Google Scholar] [CrossRef]
  9. Elashmawy, M. Experimental study on water extraction from atmospheric air using tubular solar still. J. Clean. Prod. 2020, 249, 119322. [Google Scholar] [CrossRef]
  10. Parker, A.R.; Lawrence, C.R. Water capture by a desert beetle. Nature 2001, 414, 33–34. [Google Scholar] [CrossRef]
  11. Ju, J.; Bai, H.; Zheng, Y.; Zhao, T.; Fang, R.; Jiang, L. A multi-structural and multi-functional integrated fog collection system in cactus. Nat. Commun. 2012, 3, 1247. [Google Scholar] [CrossRef] [Green Version]
  12. Zheng, Y.; Bai, H.; Huang, Z.; Tian, X.; Nie, F.; Zhao, Y.; Zhai, J.; Jiang, L. Directional water collection on wetted spider silk. Nature 2010, 463, 640–643. [Google Scholar] [CrossRef]
  13. Spellman, F.R. The Handbook of Meteorology; Scarecrow Press: Lanham, MD, USA, 2012. [Google Scholar]
  14. Schemenauer, R.S.; Cereceda, P. Water from fog-covered mountains. Waterlines 1992, 10, 10–13. [Google Scholar] [CrossRef]
  15. Klemm, O.; Schemenauer, R.; Lummerich, A.; Cereceda, P.; Marzol, V.; Corell, D.; van Heerden, J.; Reinhard, D.; Gherezghiher, T.; Olivier, J.; et al. Fog as a fresh-water resource: Overview and perspectives. Ambio 2012, 41, 221–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. World Health Organization. Guidelines for Drinking-Water Quality: Fourth Edition Incorporating First Addendum; World Health Organization: Geneva, Switzerland, 2017. [Google Scholar]
  17. Schemenauer, R.S.; Fuenzalida, H.; Cereceda, P. A Neglected Water Resource:The Camanchaca of South America. Bull. Am. Meteorol. Soc. 1988, 69, 138–147. [Google Scholar] [CrossRef] [Green Version]
  18. Schemenauer, R.S.; Cereceda, P. Fog-Water Collection in Arid Coastal Locations. Ambio 1991, 20, 303–308. [Google Scholar]
  19. Schemenauer, R.; Cereceda, P. A proposed standard fog collector for use in high-elevation regions. J. Appl. Meteorol. Climatol. 1994, 33, 1313–1322. [Google Scholar] [CrossRef]
  20. Olivier, J.; de Rautenbach, C. The implementation of fog water collection systems in South Africa. Atmos. Res. 2002, 64, 227–238. [Google Scholar] [CrossRef]
  21. Shanyengana, E.; Henschel, J.; Seely, M.; Sanderson, R. Exploring fog as a supplementary watersource in Namibia. Atmos. Res. 2002, 64, 251–259. [Google Scholar] [CrossRef]
  22. Wang, B.; Zhang, Y.; Liang, W.; Wang, G.; Guo, Z.; Liu, W. A simple route to transform normal hydrophilic cloth into a superhydrophobic–superhydrophilichybrid surface. J. Mater. Chem. A 2014, 2, 7845–7852. [Google Scholar] [CrossRef]
  23. Cao, M.; Ju, J.; Li, K.; Dou, S.; Liu, K.; Jiang, L. Facile and large-scale fabrication of a cactus-inspired continuous fog collector. Adv. Funct. Mater. 2014, 24, 3235–3240. [Google Scholar] [CrossRef]
  24. Wong, T.S.; Kang, S.H.; Tang, S.K.Y.; Smythe, E.J.; Hatton, B.D.; Grinthal, A.; Aizenberg, J. Bioinspired self-repairing slippery surfaces with pressure-stable omniphobicity. Nature 2011, 477, 443–447. [Google Scholar] [CrossRef]
  25. Feng, R.; Xu, C.; Song, F.; Wang, F.; Wang, X.; Wang, Y. A Bioinspired Slippery Surface with Stable Lubricant Impregnationfor Efficient Water Harvesting. ACS Appl. Mater. Interfaces 2020, 12, 12373–12381. [Google Scholar] [CrossRef] [PubMed]
  26. Han, X.; Guo, Z. Lubricant-Infused Three-Dimensional Frame Composed of a Micro/Nanospinous Ball Cluster Structure with Salient Durability andSuperior Fog Harvesting Capacity. ACS Appl. Mater. Interfaces 2021, 13, 46192–46201. [Google Scholar] [CrossRef] [PubMed]
  27. Andrews, H.; Eccles, E.; Schofield, W.; Badyal, J. Three-dimensional hierarchical structures for fog harvesting. Langmuir 2011, 27, 3798–3802. [Google Scholar] [CrossRef] [PubMed]
  28. Lee, S.J.; Ha, N.; Kim, H. Superhydrophilic–Superhydrophobic Water Harvester Inspired by Wetting Property of Cactus Stem. ACS Sustain. Chem. Eng. 2019, 7, 10561–10569. [Google Scholar] [CrossRef]
  29. Yu, Z.; Li, S.; Liu, M.; Zhu, R.; Yu, M.; Dong, X.; Sun, Y.; Fu, S. A dual-biomimetic knitted fabric with a manipulable structure and wettability for highly efficient fog harvesting. J. Mater. Chem. A 2022, 10, 304–312. [Google Scholar] [CrossRef]
  30. Wang, Q.; He, Y.; Geng, X.; Hou, Y.; Zheng, Y. Enhanced Fog Harvesting through Capillary-Assisted Rapid Transport of Droplet Confined in the Given Microchannel. ACS Appl. Mater. Interfaces 2021, 13, 48292–48300. [Google Scholar] [CrossRef]
  31. Feng, R.; Song, F.; Xu, C.; Wang, X.; Wang, Y. A Quadruple-Biomimetic surface for spontaneous and efficientfog harvesting. Chem. Eng. J. 2021, 422, 130119. [Google Scholar] [CrossRef]
  32. Shi, W.; Anderson, M.J.; Tulkoff, J.B.; Kennedy, B.S.; Boreyko, J.B. Fog harvesting with harps. ACS Appl. Mater. Interfaces 2018, 10, 11979–11986. [Google Scholar] [CrossRef]
  33. Cheng, Y.; Zhang, S.; Liu, S.; Huang, J.; Zhang, Z.; Wang, X.; Yu, Z.; Li, S.; Chen, Z.; Zhao, Y.; et al. Fog catcher brushes with environmental friendly slippery aluminamicro-needle structured surface for efficient fog-harvesting. J. Clean. Prod. 2021, 315, 127862. [Google Scholar] [CrossRef]
  34. Beysens, D. The formation of dew. Atmos. Res. 1995, 39, 215–237. [Google Scholar] [CrossRef]
  35. Nioras, D.; Ellinas, K.; Constantoudis, V.; Gogolides, E. How Different Are Fog Collection and Dew Water Harvesting on Surfaces with Different Wetting Behaviors? ACS Appl. Mater. Interfaces 2021, 13, 48322–48332. [Google Scholar] [CrossRef] [PubMed]
  36. Jiang, H.; Liu, X. Principles of Mimicking and Engineering the Self-organized Structure of Hard Tissues. J. Biol. Chem. 2004, 270, 41286–41293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Jumikis, A.R. Aerial Wells: Secondary Sources of Water. Soil Sci. 1965, 100, 83–95. [Google Scholar] [CrossRef]
  38. Beysens, D.; Muselli, M.; Milimouk, I.; Ohayon, C.; Berkowicz, S.M.; Soyeux, E.; Mileta, M.; Ortega, P. Application of passive radiative cooling for dew condensation. Energy 2006, 31, 2303–2315. [Google Scholar] [CrossRef]
  39. Nilsson, T.M.J.; Vargas, W.E.; Niklasson, G.A.; Granqvist, C.G. Condensation of water by radiative cooling. Renew. Energy 1994, 5, 310–317. [Google Scholar] [CrossRef]
  40. Alnaser, W.E.; Barakat, A. Use of condensed water vapour from the atmosphere for irrigation in Bahrain. Appl. Energy 2000, 65, 3–18. [Google Scholar] [CrossRef]
  41. Zhao, B.; Hu, M.; Ao, X.; Chen, N.; Pei, G. Radiative cooling: A review of fundamentals, materials, applications, and prospects. Appl. Energy 2019, 236, 489–513. [Google Scholar] [CrossRef]
  42. Khalil, B.; Adamowski, J.; Shabbir, A.; Jang, C.; Rojas, M.; Reilly, K.; Ozga-Zielinski, B. A review: Dew water collection from radiative passive collectors to recent developments of active collectors. Sustain. Water Resour. Manag. 2016, 2, 71–86. [Google Scholar] [CrossRef] [Green Version]
  43. Varanasi, K.K.; Hsu, M.; Bhate, N.; Yang, W.; Deng, T. Spatial control in the heterogeneous nucleation of water. Appl. Phys. Lett. 2009, 95, 094101. [Google Scholar] [CrossRef]
  44. Lee, A.; Moon, M.W.; Lim, H.; Kim, W.D.; Kim, H.Y. Water harvest via dewing. Langmuir 2012, 28, 10183–10191. [Google Scholar] [CrossRef]
  45. Tracy, C.R.; Laurence, N.; Christian, K.A. Condensation onto the skin as a means for water gain by tree frogs in tropical Australia. Am. Nat. 2011, 178, 553–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Jørgensen, C.B. 200 years of amphibian water economy: From Robert Townson to the present. Biol. Rev. 1997, 72, 153–237. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, C.H.; Cai, Q.; Tsai, C.; Chen, C.L.; Xiong, G.; Yu, Y.; Ren, Z. Dropwise condensation on superhydrophobic surfaces with two-tier roughness. Appl. Phys. Lett. 2007, 90, 173108. [Google Scholar] [CrossRef]
  48. Enright, R.; Miljkovic, N.; Al-Obeidi, A.; Thompson, C.V.; Wang, E.N. Condensation on superhydrophobic surfaces: The role of local energy barriers and structure length scale. Langmuir 2012, 28, 14424–14432. [Google Scholar] [CrossRef] [PubMed]
  49. Seo, D.; Lee, C.; Nam, Y. Influence of geometric patterns of microstructured superhydrophobic surfaces on water-harvesting performance via dewing. Langmuir 2014, 30, 15468–15476. [Google Scholar] [CrossRef] [PubMed]
  50. Hou, Y.; Shang, Y.; Yu, M.; Feng, C.; Yu, H.; Yao, S. Tunable Water Harvesting Surfaces Consisting of Biphilic Nanoscale Topography. ACS Nano 2018, 12, 11022–11030. [Google Scholar] [CrossRef]
  51. Hou, K.; Li, X.; Li, Q.; Chen, X. Tunable Wetting Patterns on Superhydrophilic/Superhydrophobic Hybrid Surfaces for Enhanced Dew-Harvesting Efficacy. Adv. Mater. Interfaces 2020, 7, 1901683. [Google Scholar] [CrossRef]
  52. USGS. Generalized World Precipitation Map. Available online: https://www.usgs.gov/media/images/generalized-world-precipitation-map (accessed on 6 February 2022).
  53. Yannopoulos, S.; Antoniou, G.; Kaiafa-Saropoulou, M.; Angelakis, A.N. Historical development of rainwater harvesting and use in Hellas: A preliminary review. Water Sci. Technol. Water Supply 2017, 17, 1022–1034. [Google Scholar] [CrossRef]
  54. Campisano, A.; Butler, D.; Ward, S.; Burns, M.J.; Friedler, E.; DeBusk, K.; Fisher-Jeffes, L.N.; Ghisi, E.; Rahman, A.; Furumai, H.; et al. Urban rainwater harvesting systems: Research, implementation and future perspectives. Water Res. 2017, 115, 195–209. [Google Scholar] [CrossRef]
  55. Abdulla, F.A.; Al-Shareef, A.W. Roof rainwater harvesting systems for household water supply in Jordan. Desalination 2009, 243, 195–207. [Google Scholar] [CrossRef]
  56. Melville-Shreeve, P.; Ward, S.; Butler, D. Rainwater Harvesting Typologies for UK Houses: A Multi Criteria Analysis of System Configurations. Water 2016, 8, 129. [Google Scholar] [CrossRef] [Green Version]
  57. Biazin, B.; Sterk, G.; Temesgen, M.; Abdulkedir, A.; Stroosnijder, L. Rainwater harvesting and management in rainfed agricultural systemsin sub-Saharan Africa—A review. Phys. Chem. Earth Parts A/B/C 2012, 47, 139–151. [Google Scholar] [CrossRef]
  58. Xiao, G.; Zhang, Q.; Xiong, Y.; Lin, M.; Wang, J. Integrating rainwater harvesting with supplemental irrigationinto rain-fed spring wheat farming. Soil Tillage Res. 2007, 93, 429–437. [Google Scholar] [CrossRef]
  59. Bruins, H.J.; Evenari, M.; Nessler, U. Rainwater-harvesting agriculture for food production in arid zones: The challenge of the African famine. Appl. Geogr. 1986, 6, 13–32. [Google Scholar] [CrossRef]
  60. Pari, L.; Suardi, A.; Stefanoni, W.; Latterini, F.; Palmieri, N. Economic and Environmental Assessment of Two Different Rain Water Harvesting Systems for Agriculture. Sustainability 2021, 13, 3871. [Google Scholar] [CrossRef]
  61. Alim, M.A.; Ashraf, A.A.; Rahman, A.; Tao, Z.; Roy, R.; Khan, M.M.; Shirin, S. Experimental investigation of an integrated rainwater harvesting unit for drinking water production at the household level. J. Water Process Eng. 2021, 44, 102318. [Google Scholar] [CrossRef]
  62. Zevenbergen, C.; Fu, D.; Pathirana, A. Transitioning to Sponge Cities: Challenges and Opportunities to Address Urban Water Problems in China. Water 2018, 10, 1230. [Google Scholar] [CrossRef] [Green Version]
  63. Adham, A.; Sayl, K.N.; Abed, R.; Abdeladhim, M.A.; Wesseling, J.G.; Riksen, M.; Fleskens, L.; Karim, U.; Ritsema, C.J. A GIS-based approach for identifying potential sites for harvesting rainwater in the Western Desert of Iraq. Int. Soil Water Conserv. Res. 2018, 6, 297–304. [Google Scholar] [CrossRef]
  64. Fritzmann, C.; Löwenberg, J.; Wintgens, T.; Melin, T. State-of-the-art of reverse osmosis desalination. Desalination 2007, 216, 1–76. [Google Scholar] [CrossRef]
  65. USGS. The Distribution of Water on, in, and above the Earth. Available online: https://www.usgs.gov/media/images/distribution-water-and-above-earth (accessed on 14 February 2022).
  66. Al-Karaghouli, A.; Kazmerski, L.L. Energy Consumption and Water Production Cost of Conventional and Renewable-Energy-Powered Desalination Processes. Renew. Sustain. Energy Rev. 2013, 24, 343–356. [Google Scholar] [CrossRef]
  67. Qiblawey, H.M.; Banat, F. Solar thermal desalination technologies. Desalination 2008, 220, 633–644. [Google Scholar] [CrossRef]
  68. Lei, Z.; Sun, X.; Zhu, S.; Dong, K.; Liu, X.; Wang, L.; Zhang, X.; Qu, L.; Zhang, X. Nature Inspired MXene-Decorated 3D Honeycomb-Fabric Architectures Toward Efficient Water Desalination and Salt Harvesting. Nano-Micro Lett. 2022, 14, 1–16. [Google Scholar] [CrossRef] [PubMed]
  69. Alkhudhiri, A.; Darwish, N.; Hilal, N. Membrane distillation: A comprehensive review. Desalination 2012, 287, 2–18. [Google Scholar] [CrossRef]
  70. González, D.; Amigo, J.; Suárez, F. Membrane distillation: Perspectives for sustainable and improved desalination. Renew. Sustain. Energy Rev. 2017, 80, 238–259. [Google Scholar] [CrossRef]
  71. AlMarzooqi, F.A.; AlGhaferi, A.A.; Saadat, I.; Hilal, N. Application of Capacitive Deionisation in water desalination: A review. Desalination 2014, 342, 3–15. [Google Scholar] [CrossRef]
  72. Elimelech, M.; Phillip, W.A. The future of seawater desalination: Energy, technology, and the environment. Science 2011, 333, 712–717. [Google Scholar] [CrossRef]
  73. Al-Amshawee, S.; Yunus, M.Y.B.M.; Azoddein, A.A.M.; Hassell, D.G.; Dakhil, I.H.; Hasan, H.A. Electrodialysis desalination for water and wastewater: A review. Chem. Eng. J. 2020, 380, 122231. [Google Scholar] [CrossRef]
  74. Gujjala, L.K.S.; Dutta, D.; Sharma, P.; Kundu, D.; Vo, D.N.; Kumar, S. A state-of-the-art review on microbial desalination cells. Chemosphere 2022, 288, 132386. [Google Scholar] [CrossRef]
  75. Akther, N.; Sodiq, A.; Giwa, A.; Daer, S.; Arafat, H.A.; Hasan, S.W. Recent advancements in forward osmosis desalination: A review. Chem. Eng. J. 2015, 281, 502–522. [Google Scholar] [CrossRef]
  76. Yuan, H.; Abu-Reesh, I.M.; He, Z. Enhancing desalination and wastewater treatment by coupling microbial desalination cells with forward osmosis. Chem. Eng. J. 2015, 270, 437–443. [Google Scholar] [CrossRef]
  77. Guo, J.; Tucker, Z.D.; Wang, Y.; Ashfeld, B.L.; Luo, T. Ionic liquid enables highly efficient low temperature desalination by directional solvent extraction. Nat. Commun. 2021, 12, 437. [Google Scholar] [CrossRef]
  78. Zhang, Y.; Palomba, V.; Frazzica, A. Development and characterization of LiCl supported composite sorbents for adsorption desalination. Appl. Therm. Eng. 2022, 203, 117953. [Google Scholar] [CrossRef]
  79. Alayli, Y.; Hadji, N.E.; Leblond, J. A new process for the extraction of water from air. Desalination 1987, 67, 227–229. [Google Scholar] [CrossRef]
  80. Hanikel, N.; Prévot, M.S.; Yaghi, O.M. MOF water harvesters. Nat. Nanotechnol. 2020, 15, 348–355. [Google Scholar] [CrossRef] [PubMed]
  81. Essa, F.A.; Elsheikh, A.H.; Sathyamurthy, R.; Manokar, A.M.; Kandeal, A.W.; Shanmugan, S.; Kabeel, A.E.; Sharshir, S.W.; Panchal, H.; Younes, M.M. Extracting water content from the ambient air in a double-slope half-cylindrical basin solar still using silica gel under Egyptian conditions. Sustain. Energy Technol. Assess. 2020, 39, 100712. [Google Scholar] [CrossRef]
  82. Furukawa, H.; Gándara, F.; Zhang, Y.; Jiang, J.; Queen, W.L.; Hudson, M.R.; Yaghi, O.M. Water Adsorption in Porous Metal-Organic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, 136, 4369–4381. [Google Scholar] [CrossRef]
  83. Fathieh, F.; Kalmutzki, M.J.; Kapustin, E.A.; Waller, P.J.; Yang, J.; Yaghi, O.M. Practical water production from desert air. Sci. Adv. 2018, 4, eaat3198. [Google Scholar] [CrossRef] [Green Version]
  84. Hanikel, N.; Prévot, M.S.; Fathieh, F.; Kapustin, E.A.; Lyu, H.; Wang, H.; Diercks, N.J.; Glover, T.G.; Yaghi, O.M. Rapid Cycling and Exceptional Yield in a Metal-Organic Framework Water Harvester. ACS Cent. Sci. 2019, 5, 1699–1706. [Google Scholar] [CrossRef]
Figure 1. The locations where fog harvesting is achievable for producing fresh water. Reproduced with permission from Klemm, Otto, et al., Fog as a fresh-water resource: Overview and perspectives; published by Springer Nature, 2012 [15].
Figure 1. The locations where fog harvesting is achievable for producing fresh water. Reproduced with permission from Klemm, Otto, et al., Fog as a fresh-water resource: Overview and perspectives; published by Springer Nature, 2012 [15].
Sustainability 14 06244 g001
Figure 2. The traditional and bio-inspired fog harvesting materials. (a) The 1 m 2 SFC assembled on the ridge at Cerro Orara, Peru. Reproduced with permission from Schemenauer, Robert S., et al., A proposed standard fog collector for use in high-elevation regions; published by American Meteorological Society, 1994. (b) The superhydrophobic-superhydrophilic hybrid (SSH) fabrics inspired by desert beetles. Reproduced with permission from Wang, Ben, et al., A simple route to transform normal hydrophilic cloth into a superhydrophobic–superhydrophilic hybrid surface; published by Royal Society of Chemistry, 2013. (c) The cactus-inspired fog collector. Reproduced with permission from Cao, Moyuan, et al., Facile and Large-Scale Fabrication of a Cactus-Inspired Continuous Fog Collector; published by John Wiley and Sons, 2014. (d) The schematic view of spider silk structures. Reproduced with permission from Zheng, Yongmei, et al., Directional water collection on wetted spider silk; published by Springer Nature, 2010 [12,19,22,23].
Figure 2. The traditional and bio-inspired fog harvesting materials. (a) The 1 m 2 SFC assembled on the ridge at Cerro Orara, Peru. Reproduced with permission from Schemenauer, Robert S., et al., A proposed standard fog collector for use in high-elevation regions; published by American Meteorological Society, 1994. (b) The superhydrophobic-superhydrophilic hybrid (SSH) fabrics inspired by desert beetles. Reproduced with permission from Wang, Ben, et al., A simple route to transform normal hydrophilic cloth into a superhydrophobic–superhydrophilic hybrid surface; published by Royal Society of Chemistry, 2013. (c) The cactus-inspired fog collector. Reproduced with permission from Cao, Moyuan, et al., Facile and Large-Scale Fabrication of a Cactus-Inspired Continuous Fog Collector; published by John Wiley and Sons, 2014. (d) The schematic view of spider silk structures. Reproduced with permission from Zheng, Yongmei, et al., Directional water collection on wetted spider silk; published by Springer Nature, 2010 [12,19,22,23].
Sustainability 14 06244 g002
Figure 3. The fabrication process of (a) MNS-SLIPS and (b) lubricate-infused three-dimensional frame composed of nanospinous ball cluster structures. Reproduced with permission from Feng, Rui, et al., A Bioinspired Slippery Surface with Stable Lubricant Impregnation for Efficient Water Harvesting; published by American Chemical Society, 2020. Reproduced with permission from Han, Xutong, et al., Lubricant-Infused Three-Dimensional Frame Composed of a Micro/Nanospinous Ball Cluster Structure with Salient Durability and Superior Fog Harvesting Capacity; published by American Chemical Society, 2021 [25,26].
Figure 3. The fabrication process of (a) MNS-SLIPS and (b) lubricate-infused three-dimensional frame composed of nanospinous ball cluster structures. Reproduced with permission from Feng, Rui, et al., A Bioinspired Slippery Surface with Stable Lubricant Impregnation for Efficient Water Harvesting; published by American Chemical Society, 2020. Reproduced with permission from Han, Xutong, et al., Lubricant-Infused Three-Dimensional Frame Composed of a Micro/Nanospinous Ball Cluster Structure with Salient Durability and Superior Fog Harvesting Capacity; published by American Chemical Society, 2021 [25,26].
Sustainability 14 06244 g003
Figure 4. (a) The 3D hierarchical structure of Cotula fallax and (b) cactus-inspired superhydrophilic-superhydrophobic water harvester integrated with PNIPAAm. Reproduced with permission from Andrews, H. G., et al., Three-Dimensional Hierarchical Structures for Fog Harvesting; published by American Chemical Society, 2011. Reproduced with permission from Lee, Sang Joon, et al., Superhydrophilic–Superhydrophobic Water Harvester Inspired by Wetting Property of Cactus Stem; published by American Chemical Society, 2019 [27,28].
Figure 4. (a) The 3D hierarchical structure of Cotula fallax and (b) cactus-inspired superhydrophilic-superhydrophobic water harvester integrated with PNIPAAm. Reproduced with permission from Andrews, H. G., et al., Three-Dimensional Hierarchical Structures for Fog Harvesting; published by American Chemical Society, 2011. Reproduced with permission from Lee, Sang Joon, et al., Superhydrophilic–Superhydrophobic Water Harvester Inspired by Wetting Property of Cactus Stem; published by American Chemical Society, 2019 [27,28].
Sustainability 14 06244 g004
Figure 5. The design of (a) SHL–SHB@TF, (b) IBS by the capillary-induced selective oxidation method and (c) SLIPS-SHIS. Reproduced with permission from Yu, Zhihua, et al., A dual-biomimetic knitted fabric with a manipulable structure and wettability for highly efficient fog harvesting; published by the Royal Society of Chemistry, 2022. Reproduced with permission from Wang, Qianqian, et al., Enhanced Fog Harvesting through Capillary-Assisted Rapid Transport of Droplet Confined in the Given Microchannel; published by American Chemical Society, 2021. Reproduced with permission from Feng, Rui, et al., A Quadruple-Biomimetic surface for spontaneous and efficient fog harvesting; published by Elsevier, 2021 [29,30,31].
Figure 5. The design of (a) SHL–SHB@TF, (b) IBS by the capillary-induced selective oxidation method and (c) SLIPS-SHIS. Reproduced with permission from Yu, Zhihua, et al., A dual-biomimetic knitted fabric with a manipulable structure and wettability for highly efficient fog harvesting; published by the Royal Society of Chemistry, 2022. Reproduced with permission from Wang, Qianqian, et al., Enhanced Fog Harvesting through Capillary-Assisted Rapid Transport of Droplet Confined in the Given Microchannel; published by American Chemical Society, 2021. Reproduced with permission from Feng, Rui, et al., A Quadruple-Biomimetic surface for spontaneous and efficient fog harvesting; published by Elsevier, 2021 [29,30,31].
Sustainability 14 06244 g005
Figure 6. The design of (a) fog harps and (b) fog catcher brushes. Reproduced with permission from Shi, Weiwei, et al., Fog harvesting with harps; published by American Chemical Society, 2018. Reproduced with permission from Cheng, Yan, et al., Fog catcher brushes with environmental friendly slippery alumina micro-needle structured surface for efficient fog-harvesting; published by Elsevier, 2021 [32,33].
Figure 6. The design of (a) fog harps and (b) fog catcher brushes. Reproduced with permission from Shi, Weiwei, et al., Fog harvesting with harps; published by American Chemical Society, 2018. Reproduced with permission from Cheng, Yan, et al., Fog catcher brushes with environmental friendly slippery alumina micro-needle structured surface for efficient fog-harvesting; published by Elsevier, 2021 [32,33].
Sustainability 14 06244 g006
Figure 7. The illustration of (a) homogeneous nucleation and (b) heterogeneous nucleation. Reproduced with permission from Jiang, Huaidong, et al., Principles of Mimicking and Engineering the Self-organized Structure of Hard Tissues; published by Elsevier under the terms of the CC BY 4.0 license, 2004 [36].
Figure 7. The illustration of (a) homogeneous nucleation and (b) heterogeneous nucleation. Reproduced with permission from Jiang, Huaidong, et al., Principles of Mimicking and Engineering the Self-organized Structure of Hard Tissues; published by Elsevier under the terms of the CC BY 4.0 license, 2004 [36].
Sustainability 14 06244 g007
Figure 8. The dew condensation process over time (ah) on the surface with alternating hydrophobic and hydrophilic regions. Reproduced with permission from Varanasi, Kripa K., et al., Spatial control in the heterogeneous nucleation of water; published by AIP Publishing, 2009 [43].
Figure 8. The dew condensation process over time (ah) on the surface with alternating hydrophobic and hydrophilic regions. Reproduced with permission from Varanasi, Kripa K., et al., Spatial control in the heterogeneous nucleation of water; published by AIP Publishing, 2009 [43].
Sustainability 14 06244 g008
Figure 9. (a) The illustration of biphilic structured surfaces composed of nanoscale hydrophilic bumps and superhydrophobic substrates and (b) the scheme of fabrication of superhydrophilic/superhydrohobic hybrid surfaces. Reproduced with permission from Hou, Youmin, et al., Tunable Water Harvesting Surfaces Consisting of Biphilic Nanoscale Topography; published by American Chemical Society, 2018. Reproduced with permission from Chen, Xuemei, et al., Tunable Wetting Patterns on Superhydrophilic/Superhydrophobic Hybrid Surfaces for Enhanced Dew-Harvesting Efficacy; published by John Wiley and Sons, 2019 [50,51].
Figure 9. (a) The illustration of biphilic structured surfaces composed of nanoscale hydrophilic bumps and superhydrophobic substrates and (b) the scheme of fabrication of superhydrophilic/superhydrohobic hybrid surfaces. Reproduced with permission from Hou, Youmin, et al., Tunable Water Harvesting Surfaces Consisting of Biphilic Nanoscale Topography; published by American Chemical Society, 2018. Reproduced with permission from Chen, Xuemei, et al., Tunable Wetting Patterns on Superhydrophilic/Superhydrophobic Hybrid Surfaces for Enhanced Dew-Harvesting Efficacy; published by John Wiley and Sons, 2019 [50,51].
Sustainability 14 06244 g009
Figure 10. The global map of (a) generalized average annual precipitation and (b) overall water scarcity risks. Reproduced from Water Science School, Generalized world precipitation map; published by U.S. Geological Survey. Reproduced with permission from Gassert, Francis, et al., Aqueduct Global Maps 2.1: Constructing Decision-Relevant Global Water Risk Indicators; published by World Resources Institute, 2014 [3,52].
Figure 10. The global map of (a) generalized average annual precipitation and (b) overall water scarcity risks. Reproduced from Water Science School, Generalized world precipitation map; published by U.S. Geological Survey. Reproduced with permission from Gassert, Francis, et al., Aqueduct Global Maps 2.1: Constructing Decision-Relevant Global Water Risk Indicators; published by World Resources Institute, 2014 [3,52].
Sustainability 14 06244 g010
Figure 11. The schematic view of (a) the typical rainwater harvesting system in Jordan and (b) an innovative rainwater harvesting system in Australia, Korea and the USA. Reproduced with permission from Abdulla, Fayez A., et al., Roof rainwater harvesting systems for household water supply in Jordan; published by Elsevier, 2009. Reproduced with permission from Melville-Shreeve, Peter, et al., Rainwater Harvesting Typologies for UK Houses: A Multi Criteria Analysis of System Configurations; published by MDPI under the terms of the CC BY 4.0 license, 2016 [55,56].
Figure 11. The schematic view of (a) the typical rainwater harvesting system in Jordan and (b) an innovative rainwater harvesting system in Australia, Korea and the USA. Reproduced with permission from Abdulla, Fayez A., et al., Roof rainwater harvesting systems for household water supply in Jordan; published by Elsevier, 2009. Reproduced with permission from Melville-Shreeve, Peter, et al., Rainwater Harvesting Typologies for UK Houses: A Multi Criteria Analysis of System Configurations; published by MDPI under the terms of the CC BY 4.0 license, 2016 [55,56].
Sustainability 14 06244 g011
Figure 12. The illustration of (a) the field cultivation of rainwater harvesting with a sowing in the furrow between film-covered ridges (SFFCR), (b) the field cultivation of rainwater harvesting with a sowing in the holes on film-covered ridges (SHFCR), (c) the liman terrace and (d) the innovative flexible water storage system (FWSS). Reproduced with permission from Xiao, Guoju, et al., Integrating rainwater harvesting with supplemental irrigation into rain-fed spring wheat farming; published by Elsevier, 2007. Reproduced with permission from Bruins, H. J., et al., Rainwater-harvesting agriculture for food production in arid zones: the challenge of the African famine; published by Elsevier, 1986. Reproduced with permission from Pari, Luigi, et al., Economic and Environmental Assessment of Two Different Rain Water Harvesting Systems for Agriculture; published by MDPI under the terms of the CC BY 4.0 license, 2021 [58,59,60].
Figure 12. The illustration of (a) the field cultivation of rainwater harvesting with a sowing in the furrow between film-covered ridges (SFFCR), (b) the field cultivation of rainwater harvesting with a sowing in the holes on film-covered ridges (SHFCR), (c) the liman terrace and (d) the innovative flexible water storage system (FWSS). Reproduced with permission from Xiao, Guoju, et al., Integrating rainwater harvesting with supplemental irrigation into rain-fed spring wheat farming; published by Elsevier, 2007. Reproduced with permission from Bruins, H. J., et al., Rainwater-harvesting agriculture for food production in arid zones: the challenge of the African famine; published by Elsevier, 1986. Reproduced with permission from Pari, Luigi, et al., Economic and Environmental Assessment of Two Different Rain Water Harvesting Systems for Agriculture; published by MDPI under the terms of the CC BY 4.0 license, 2021 [58,59,60].
Sustainability 14 06244 g012
Figure 13. (a) The schematic view of filtration unit of IRWHU. Reproduced with permission from Alim, Mohammad A., et al.,Experimental investigation of an integrated rainwater harvesting unit for drinking water production at the household level; published by Elsevier, 2021. (b) The application of Sponge City Concept (SCC) in different zones of a city. Reproduced with permission from Zevenbergen, Chris, et al., Transitioning to Sponge Cities: Challenges and Opportunities to Address Urban Water Problems in China; published by MDPI under the terms of the CC BY 4.0 license, 2018 (c) The suitability map of potential rainwater harvesting sites produced by the GIS-based approach. Reproduced with permission from Adham, Ammar, et al., A GIS-based approach for identifying potential sites for harvesting rainwater in the Western Desert of Iraq; published by Elsevier under the terms of the CC BY-NC-ND 4.0 license, 2018 [61,62,63].
Figure 13. (a) The schematic view of filtration unit of IRWHU. Reproduced with permission from Alim, Mohammad A., et al.,Experimental investigation of an integrated rainwater harvesting unit for drinking water production at the household level; published by Elsevier, 2021. (b) The application of Sponge City Concept (SCC) in different zones of a city. Reproduced with permission from Zevenbergen, Chris, et al., Transitioning to Sponge Cities: Challenges and Opportunities to Address Urban Water Problems in China; published by MDPI under the terms of the CC BY 4.0 license, 2018 (c) The suitability map of potential rainwater harvesting sites produced by the GIS-based approach. Reproduced with permission from Adham, Ammar, et al., A GIS-based approach for identifying potential sites for harvesting rainwater in the Western Desert of Iraq; published by Elsevier under the terms of the CC BY-NC-ND 4.0 license, 2018 [61,62,63].
Sustainability 14 06244 g013
Figure 14. The composition of total global water on Earth. Reproduced with permission from Water Science School, The distribution of water on, in, and above the Earth; published by U.S. Geological Survey, 2019 [2,65].
Figure 14. The composition of total global water on Earth. Reproduced with permission from Water Science School, The distribution of water on, in, and above the Earth; published by U.S. Geological Survey, 2019 [2,65].
Sustainability 14 06244 g014
Figure 15. The schematic views of thermal desalination technologies. (a) Multi-stage flash (MSF); (b) mlti-effect distillation (MED). Reproduced with permission from Al-Karaghouli, Ali, et al., Energy consumption and water production cost of conventional and renewable-energy-powereddesalination processes; published by Elsevier, 2013. (c) MXene-decorated 3D honeycomb-fabric design. Reproduced with permission from Lei, Zhiwei, et al., Nature Inspired MXene-Decorated 3D Honeycomb-Fabric Architectures Toward Efficient Water Desalination and Salt Harvesting; published by Springer Nature under the terms of the CC BY 4.0 license, 2022. (d) Direct Contact Membrane Distillation (DCMD). Reproduced with permission from Alkhudhiri, Abdullah, et al., Membrane distillation: A comprehensive review; published by Elsevier, 2012 [66,68,69].
Figure 15. The schematic views of thermal desalination technologies. (a) Multi-stage flash (MSF); (b) mlti-effect distillation (MED). Reproduced with permission from Al-Karaghouli, Ali, et al., Energy consumption and water production cost of conventional and renewable-energy-powereddesalination processes; published by Elsevier, 2013. (c) MXene-decorated 3D honeycomb-fabric design. Reproduced with permission from Lei, Zhiwei, et al., Nature Inspired MXene-Decorated 3D Honeycomb-Fabric Architectures Toward Efficient Water Desalination and Salt Harvesting; published by Springer Nature under the terms of the CC BY 4.0 license, 2022. (d) Direct Contact Membrane Distillation (DCMD). Reproduced with permission from Alkhudhiri, Abdullah, et al., Membrane distillation: A comprehensive review; published by Elsevier, 2012 [66,68,69].
Sustainability 14 06244 g015
Figure 16. The distribution of capital among different desalination technologies. Reproduced with permission from AlMarzooqi, Faisal A., et al., Application of Capacitive Deionisation in water desalination: A review; published by Elsevier, 2014 [71].
Figure 16. The distribution of capital among different desalination technologies. Reproduced with permission from AlMarzooqi, Faisal A., et al., Application of Capacitive Deionisation in water desalination: A review; published by Elsevier, 2014 [71].
Sustainability 14 06244 g016
Figure 17. The schematic views of membrane-based technologies. (a) Reverse Osmosis (RO). Reproduced with permission from Elimelech, Menachem, et al., The Future of Seawater Desalination: Energy, Technology, and the Environment; published by The American Association for the Advancement of Science, 2011. (b) air cathode MDC. Reproduced with permission from Gujjala, Lohit Kumar Srinivas, et al., A state-of-the-art review on microbial desalination cells; published by Elsevier, 2022. (c) MDC-FO hybrid system. Reproduced with permission from Yuan, Heyang, et al., Enhancing desalination and wastewater treatment by coupling microbial desalination cells with forward osmosis; published by Elsevier, 2015 [72,74,76].
Figure 17. The schematic views of membrane-based technologies. (a) Reverse Osmosis (RO). Reproduced with permission from Elimelech, Menachem, et al., The Future of Seawater Desalination: Energy, Technology, and the Environment; published by The American Association for the Advancement of Science, 2011. (b) air cathode MDC. Reproduced with permission from Gujjala, Lohit Kumar Srinivas, et al., A state-of-the-art review on microbial desalination cells; published by Elsevier, 2022. (c) MDC-FO hybrid system. Reproduced with permission from Yuan, Heyang, et al., Enhancing desalination and wastewater treatment by coupling microbial desalination cells with forward osmosis; published by Elsevier, 2015 [72,74,76].
Sustainability 14 06244 g017
Figure 18. The schematic views of (a) Directional Solvent Extraction (DSE) employing ionic liquid (IL) and (b) adsorption desalination (AD). Reproduced with permission from Guo, Jiaji, et al., Ionic liquid enables highly efficient low temperature desalination by directional solvent extraction; published by Springer Nature under the terms of the CC BY 4.0 license, 2021. Reproduced with permission from Zhang, Yannan, et al., Development and characterization of LiCl supported composite sorbents for adsorption desalination; published by Elsevier, 2022 [77,78].
Figure 18. The schematic views of (a) Directional Solvent Extraction (DSE) employing ionic liquid (IL) and (b) adsorption desalination (AD). Reproduced with permission from Guo, Jiaji, et al., Ionic liquid enables highly efficient low temperature desalination by directional solvent extraction; published by Springer Nature under the terms of the CC BY 4.0 license, 2021. Reproduced with permission from Zhang, Yannan, et al., Development and characterization of LiCl supported composite sorbents for adsorption desalination; published by Elsevier, 2022 [77,78].
Sustainability 14 06244 g018
Figure 19. The illustration of (a) tubular solar stills (TSS) and (b) the double-slope half-cylindrical basin solar still (DS-HCBSS). Reproduced with permission from Elashmawy, Mohamed, Experimental study on water extraction from atmospheric air using tubular solar still; published by Elsevier, 2020. Reproduced with permission from Essa, F. A., et al., Extracting water content from the ambient air in a double-slope half-cylindrical basin solar still using silica gel under Egyptian conditions; published by Elsevier, 2020 [9,81].
Figure 19. The illustration of (a) tubular solar stills (TSS) and (b) the double-slope half-cylindrical basin solar still (DS-HCBSS). Reproduced with permission from Elashmawy, Mohamed, Experimental study on water extraction from atmospheric air using tubular solar still; published by Elsevier, 2020. Reproduced with permission from Essa, F. A., et al., Extracting water content from the ambient air in a double-slope half-cylindrical basin solar still using silica gel under Egyptian conditions; published by Elsevier, 2020 [9,81].
Sustainability 14 06244 g019
Figure 20. The illustration of different MOFs and MOF-based water harvesters. (a) MOF-801; (b) MOF-841. Reproduced with permission from Furukawa, Hiroyasu, et al., Water Adsorption in Porous Metal–Organic Frameworks and Related Materials; published by American Chemical Society, 2014. (c) One-cycle MOF-based water harvester. Reproduced with permission from Fathieh, Farhad, et al., Practical water production from desert air; published by The American Association for the Advancement of Science under the terms of the CC BY-NC 4.0 license, 2018. (d) Multi-cycle MOF-based water harvester. Reproduced with permission from Hanikel, Nikita, et al., Rapid Cycling and Exceptional Yield in a Metal-Organic Framework Water Harvester; published by American Chemical Society under the terms of the ACS AuthorChoice/Editors’ Choice license, 2019 [82,83,84].
Figure 20. The illustration of different MOFs and MOF-based water harvesters. (a) MOF-801; (b) MOF-841. Reproduced with permission from Furukawa, Hiroyasu, et al., Water Adsorption in Porous Metal–Organic Frameworks and Related Materials; published by American Chemical Society, 2014. (c) One-cycle MOF-based water harvester. Reproduced with permission from Fathieh, Farhad, et al., Practical water production from desert air; published by The American Association for the Advancement of Science under the terms of the CC BY-NC 4.0 license, 2018. (d) Multi-cycle MOF-based water harvester. Reproduced with permission from Hanikel, Nikita, et al., Rapid Cycling and Exceptional Yield in a Metal-Organic Framework Water Harvester; published by American Chemical Society under the terms of the ACS AuthorChoice/Editors’ Choice license, 2019 [82,83,84].
Sustainability 14 06244 g020
Table 1. The water collection rate and efficiency of fog and dew harvesting. Reproduced with permission from Nioras, Dimitrios, et al., How Different Are Fog Collection and Dew Water Harvesting on Surfaces with Different Wetting Behaviors? published by American Chemical Society, 2021 [35].
Table 1. The water collection rate and efficiency of fog and dew harvesting. Reproduced with permission from Nioras, Dimitrios, et al., How Different Are Fog Collection and Dew Water Harvesting on Surfaces with Different Wetting Behaviors? published by American Chemical Society, 2021 [35].
SurfaceDew WCR (mg/cm 2 /h)Dew Efficiency η dew  (%)Fog WCR (mg/cm 2 /h)Fog Efficiency η fog  (%)
Untreated PMMA15.83170815
Hydrophobic14.62983918
Short plasma-textured nano hierarchical superhydrophobic PMMA21.843101121
Long plasma-textured micro-nano hierarchical superhydrophobic PMMA--117125
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, K.; Tao, Y.; Shi, W. Recent Advances in Water Harvesting: A Review of Materials, Devices and Applications. Sustainability 2022, 14, 6244. https://0-doi-org.brum.beds.ac.uk/10.3390/su14106244

AMA Style

Chen K, Tao Y, Shi W. Recent Advances in Water Harvesting: A Review of Materials, Devices and Applications. Sustainability. 2022; 14(10):6244. https://0-doi-org.brum.beds.ac.uk/10.3390/su14106244

Chicago/Turabian Style

Chen, Kuanfu, Yujie Tao, and Weiwei Shi. 2022. "Recent Advances in Water Harvesting: A Review of Materials, Devices and Applications" Sustainability 14, no. 10: 6244. https://0-doi-org.brum.beds.ac.uk/10.3390/su14106244

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop