Next Article in Journal
Nonwoven Ion-Exchange Membranes with High Protein Binding Capacity for Bioseparations
Next Article in Special Issue
Experimental Investigation on Floating Solar-Driven Membrane Distillation Desalination Modules
Previous Article in Journal
Direct Purification of Digestate Using Ultrafiltration Membranes: Influence of Pore Size on Filtration Behavior and Fouling Characteristics
Previous Article in Special Issue
Hole-Type Spacers for More Stable Shale Gas-Produced Water Treatment by Forward Osmosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Comparative Analysis of Conventional and Emerging Technologies for Seawater Desalination: Northern Chile as A Case Study

1
Departamento de Ingeniería Química, Facultad de Ingeniería, Universidad de Santiago de Chile (USACH), Av. Libertador Bernardo O’Higgins 3363, Estación Central 9160000, Chile
2
Centro de Innovación en Ingeniería Aplicada (CIIA), Departamento de Computación e Industrias, Facultad de Ciencias de la Ingeniería, Universidad Católica del Maule (UCM), Av. San Miguel 3605, Talca 3460000, Chile
3
Gestionare Consultores, Carlos Antunez 2025 of. 608, Providencia 7500000, Chile
*
Author to whom correspondence should be addressed.
Membranes 2021, 11(3), 180; https://doi.org/10.3390/membranes11030180
Submission received: 5 February 2021 / Revised: 25 February 2021 / Accepted: 26 February 2021 / Published: 5 March 2021
(This article belongs to the Special Issue Developments in Innovative Membrane Desalination Processes)

Abstract

:
The aim of this work was to study different desalination technologies as alternatives to conventional reverse osmosis (RO) through a systematic literature review. An expert panel evaluated thermal and membrane processes considering their possible implementation at a pilot plant scale (100 m3/d of purified water) starting from seawater at 20 °C with an average salinity of 34,000 ppm. The desalination plant would be located in the Atacama Region (Chile), where the high solar radiation level justifies an off-grid installation using photovoltaic panels. We classified the collected information about conventional and emerging technologies for seawater desalination, and then an expert panel evaluated these technologies considering five categories: (1) technical characteristics, (2) scale-up potential, (3) temperature effect, (4) electrical supply options, and (5) economic viability. Further, the potential inclusion of graphene oxide and aquaporin-based biomimetic membranes in the desalinization processes was analyzed. The comparative analysis lets us conclude that nanomembranes represent a technically and economically competitive alternative versus RO membranes. Therefore, a profitable desalination process should consider nanomembranes, use of an energy recovery system, and mixed energy supply (non-conventional renewable energy + electrical network). This document presents an up-to-date overview of the impact of emerging technologies on desalinated quality water, process costs, productivity, renewable energy use, and separation efficiency.

1. Introduction

Desalination is a separation process intended to increase water availability in structurally water-deficient countries that suffer recurrent periods of drought. Recently, the International Desalination Association (IDA) [1] reported that 150 countries apply desalination, based on daily activities of more than 300 million people worldwide. Between 2016 and 2019, the number of desalination plants and the daily water production increased by 12.4% and 41.2%, respectively, proving the accelerated growth of this technology [1,2].
Saudi Arabia has the largest water-production installed capacity, with 12 Mm3/d, representing 9.81% of the worldwide capacity, followed by the United Arab Emirates, the United States of America, Spain, and China, at 7.5, 4.7, 3.6, and 3.0%, respectively. Installation of desalination plants is mostly preferred when there is no simple alternative to obtain fresh water, low-cost energy is feasible, and high standards of living allow it [3]. After World War II, the commercial exploitation of desalination focused on technologies based on thermal processes that use phase change to separate volatile solvent (water) from nonvolatile solutes (salts) [4,5]. Currently, there are two types of desalination technologies: thermal and membrane. Figure 1 shows the conventional and emerging technologies for desalination, highlighting reverse osmosis (RO) with 65% of installed capacity worldwide [1,6,7,8,9].
Nowadays, the greatest challenge for desalination processes is to lower operating and energy costs, through emerging technologies [10,11]. These emerging technologies may arise from taking advantage of externalities generated by synergies established in the search for innovation in integrated operation models [12,13]. Likewise, the incorporation of renewable energies (e.g., solar, wind, and geothermal) into desalination and integrated processes (such as RO with pressure-retarded osmosis (PRO)), has gained attention as an alternative to reduce energy costs by 50–75% of the operational costs in the conventional process [14,15,16], and 60% of the specific energy consumption (SEC) of RO [17].
In recent years, the number of publications on technoeconomic studies about desalinization processes has increased considerably. For instance, Arafat [18] related sustainability of desalination processes with technical information, concluding that current knowledge is insufficient to describe the relevance and complexity of desalination processes. Silva-Pinto and Cunha-Marques [6] evaluated the economic feasibility of different desalination and energy supply technologies, emphasizing hybrid options and the relevance of locally specific solutions. However, the authors did not illustrate their evaluation with a particular case. Proskynitopoulou and Katsoyiannis [19] reported case studies of the main desalinization technologies, highlighting energy costs and economic parameters. The authors considered desalinization as a drinking water production process applicable only in areas of water scarcity [20,21].
This article presents different desalination technologies as alternatives to conventional reverse osmosis through current state-of-the-art desalination processes considering the existing conventional and emerging technologies and a technical economic comparison between them. Further, this document offers evaluated thermal and membrane processes considering their possible implementation at a pilot plant scale (100 m3/d of purified water) starting from seawater at 20 °C and average salinity of 34,000 ppm.

1.1. Conventional and Emerging Technologies for Desalination

Desalination technologies are classified as conventional or emerging, depending on the scientific and technical development level, and their presence in the market. According to the definition proposed by Day et al. [22], emerging desalination technologies are “scientific innovations that generate incentives to make investments in the desalination process. These innovations are based on evolved technologies that improve desalination process (that is reduce energy consumption, minimize rejection and improve water quality)”. Further, a sustainability desalination industry should consider minimizing the effect of the local increase in sea salinity due to the reject stream as an important challenge.
Conventional and emerging technologies are also classified according to the type of gradient applied (pressure, electric, chemical, and thermal) and the physicochemical process involved (Table 1). Said gradient allows the separation of saline solutes from a liquid solution, through the described phenomenon.

1.1.1. Conventional Thermal Technologies

Figure 2 shows the main characteristics and types of conventional thermal desalination technologies. These processes are based on phase changes. Particularly, processes based on freezing are scarce in the market because of high investment and operation costs associated with the relevant energy demand, retention of unpleasant aromas eventually present in the feed seawater, and intensive use of refrigerants [23]. Evaporation-based processes allow obtaining a distilled, purified water, with salinity below 10 ppm. The equipment operates in stationary phase and is arranged in multiple stages in order to increase energy efficiency, profitability of the process, and avoid fouling (caused by carbonates, sulfates, silica, and other inorganic compounds) [24,25,26]. The main conventional thermal technologies are multi-stage flash distillation (MSF), multi-effect distillation (MED), and mechanical vapor compression (MVC) (see Figure 2).
The main difference between the MED and MSF processes is that while vapor is created in an MSF system through flashing, evaporation of feed water in MED is achieved through heat transfer from the steam in condenser tubes into the source water sprayed onto these tubes. This heat transfer at the same time results in vapor condensation to freshwater [29]. MVC and MED work based on similar principles. However, in MVC, the steam generated from the evaporation of new source water sprayed on the outer surface of the heat exchanger tubes is recirculated by the vapor compressor and introduced into the inner side of the same heat exchanger tubes in which it condenses to form a distillate [29]. Seawater (intake) generates more steam when sprayed on the hot side.

1.1.2. Conventional Membrane Technologies

Figure 3 shows the main characteristics and types of conventional membrane technology used for desalination. Industrial membrane processes are designed to operate continuously. However, there are transient phenomena owing to membrane fouling that force periodic cleaning routines and lead to the slow but progressive deterioration of the polymers that compose the membranes. This is an irreversible phenomenon, which ends with the replacement of the damaged membrane. For this reason, several authors investigated methods to eliminate (or avoid) fouling (and/or scaling). For example, Mangal et al. [37] investigated if antiscalants, without acid addition, can prevent calcium phosphate scaling in RO systems. The available antiscalants, tested in the study, did not provide acceptable inhibition of calcium phosphate scaling in RO applications. Landsman et al. [38] investigated the use of a hybrid electrodialysis–nanofiltration/reverse osmosis (ED-NF/RO) system to reduce fouling from calcite precipitation and calcium polysaccharide sorption to NF/RO membranes. ED pretreatment reduced calcite oversaturation and reduced flux decline during NF/RO. Low alginate concentrations (25 mg/L) limited NF/RO fouling, but high concentrations (100 mg/L) appeared to promote calcite scaling. Dhakal et al. [39] developed and demonstrated the applicability of the flow cytometry (FCM)-based bacterial growth potential (BGP) method to assess the biofouling potential in seawater (SW) RO systems using a natural microbial consortium. Sperle et al. [40] reported the potential of UVC irradiation using the recently developed UV-LEDs as an in situ pretreatment strategy for biofouling control in RO or NF systems. In contrast to UV studies carried out previously, they tested if low fluences are sufficient to not only delay the biofilm formation but further lead to a reduced hydraulic resistance of the biofilm while approaching a severe biofouling state. On the other hand, the development of ceramic membranes for membrane distillation desalination is developing, gradually replacing their polymeric counterparts due to superior properties in terms of thermal, chemical, and mechanical stabilities, as well as potentially longer service terms [41]. Bandar et al. [42] used economically and eco-friendly Saudi red clay, tetraethyl orthosilicate, ammonia, and sodium alginate powder as a binder to fabricate a ceramic membrane for membrane distillation using an extrusion technique. The prepared membrane was tested using a vacuum membrane distillation process and showed promising permeate flux and salt rejection results.
RO is the most energy-efficient technology for desalination, with much lower energy consumption than other technologies (SEC < 3.1 kWh/m3) [43,44,45]. Typical installed capacities fluctuate between 1000 and 600,000 m3/d [19,46]. External pressure required in an RO process must exceed the osmotic pressure of the aqueous feed. Since membranes offer high salt rejection levels (>99.5%), in sea water desalination operations, pressure may even double the feed osmotic pressure [47]. On the other hand, comparing nanofiltration (NF) with RO, it arises that NF technology allows partial desalination of monovalent salts such as NaCl and KCl (up to 50–60% rejection) and efficiently removes divalent ions such as sulfates and carbonates. For this reason, NF operates at lower pressures than RO due to the lower osmotic pressure gradient between the feed flow and permeate flow. Reprocessing the permeate or using consecutive desalination NF-RO stages improves the selectivity of the process. Typical NF installed capacities are found between 1000 and 100,000 m3/d [48,49].
Electrodialysis (ED) is a membrane-based desalination technology that operates due to an electric gradient, used for many years on brackish water (salinity < 3000 mg/L). Typical installed capacities range between 5000 and 425,000 m3/d [50,51,52].

1.1.3. Innovations in Thermal Desalination

The main innovation in emerging desalination technologies relates to the use of renewable energy sources coupled to desalination processes. The most popular renewable energies included in these processes are solar and geothermal energy. For example, Prajapati et al. [64] investigated renewable energy sources that could be used to run desalination systems, and the potential matches between the desalination and renewable energy sources, to survey water desalination by utilizing geothermal and solar energy, to assess or identify areas which require improvement in geothermal and solar energy-driven desalination systems.

Solar Desalination (SD)

SD operates with direct or indirect solar energy [65]. Direct solar energy corresponds to the use of solar collectors that evaporate water and produce distillates. Indirect solar energy refers to the design of a desalination plant that uses two sub-systems: a solar collector (thermal or photovoltaic) and a desalination unit (e.g., RO).
Figure 4 shows an overview of solar energy capture. Solar collector technology for thermal distillation processes depends on the maximum temperature level (Tm) in the evaporation–condensation phenomena, i.e., Tm < 130 °C, low-temperature solar energy (LTSE) and Tm > 130 °C, high-temperature solar energy (HTSE). LTSE is generated in non-concentrated or low-concentration collectors and uses simple solar fields, i.e., without moving parts and with low investment and low operating costs [66,67]. HTSEs use mirrors to focus solar irradiation from large surfaces (aperture area) onto the small surface of the receiver. This reduces hot surface areas in solar fields. On the other hand, the integration of heat transfer fluid implies that the pipes should be thermally isolated. Subsequently, desalination technologies that utilize thermal energy and thus require thermal energy storage (TES) for uninterrupted process operation are MED, MSF, low-temperature MED, humidification–dehumidification, low-temperature desalination, and membrane distillation [68]. The most commonly used TES methods are thermal oil, pressurized water, molten salts, and storage heaters. These systems have maximum operating temperatures of 395, 150, 250–550, and 550 °C, respectively [68,69,70].
Figure 5a depicts a desalination plant that uses a solar thermal energy concentration device (CSP) and RO (CSP/RO). The CSP captures thermal energy to produce superheated steam. Some CSPs allow storing heat in a sub-process based on molten salts [71,72]. Superheated steam flows to a turbine to produce electrical energy. The high-pressure pump operates on the generated electrical energy. Figure 5b represents a desalination plant that uses photovoltaic cells and RO (PV/RO). This process is useful for small desalination equipment (<0.2 L/s) and brackish water that requires less pumping power [73,74,75,76]. PV/RO requires batteries to maintain a continuous operation [77,78].
It follows that the indirect systems are more efficient and therefore are more adequate for use in industrial scale production.

Geothermal Desalination (GD)

Geothermal energy extraction has the advantage of being independent of the season and climate changes. Geothermal energy sources (GESs) can be applied in both membrane and thermal desalination processes, depending on the location, and on the physical and chemical characteristics of geothermal water. GD recovers heat from a GES to evaporate seawater.
Membrane processes integrated with GES are still under development (see Figure 6). Gude [79] investigated the potential of geothermal energy sources for MD. They concluded that roughly a 6.1% increase in permeate flow rate for every 2-degree temperature difference (in feed water) can be achieved by utilizing process waste heat sources. Assad et al. [80] investigated two technologies for water desalination using geothermal-powered systems that are presented and discussed. These technologies are promising, especially in the Gulf region, where geothermal energy is widely available.

1.1.4. Innovations in Membrane Processes

The most relevant innovations refer to the synthesis or modification of polymeric materials that allow increasing the productivity and selectivity of desalination processes. Similarly, innovations in conventional technologies were developed, combining mass transport mechanisms, separation gradients, and renewable energy sources. In all cases, the objective is to obtain a reliable desalinated water production technology at the lowest cost.

Nanomembranes (NMs)

NMs are membranes that contain nanoparticles (zeolitic type or metal oxide) in the active layer of the polymer matrix, e.g., polymerized polyamide, aiming at improving hydrophilicity, productivity, and salt rejection [82].
NMs, also known as thin film nanocomposite membranes, are prepared by the phase inversion method (inorganic–organic mixed matrix). The common nanomaterial used in NM synthesis is TiO2 due to high chemical stability, low toxicity, defouling, and photocatalytic properties and availability [83]. The organic NM portion allows diverse geometries due to its flexibility, high density of spiral wound packing, ease of manufacturing, and good permeability and selectivity. In turn, the inorganic part of the NMs allows high surface charge density, negative zeta potential, which minimizes fouling problems, ion exchange capacity, which increases rejection, high hydrophilicity, which increases permeability, salt selectivity, and biocidal and antimicrobial capacity, which reduces bio-fouling.
Recently, some authors compared NMs with conventional reverse osmosis membranes at the same pressure, temperature, salinity, and flow rate, resulting in higher productivity (10–20%), constant selectivity (99.5–99.8%), a fouling decrease, and lower energy demand [84,85,86] (see Figure 7).

Membrane Distillation (MD)

MD uses hydrophobic polymeric membranes (e.g., PP, PTFE, PVDF) of porosities between 0.01 and 0.5 μm. The vapor distillate in MD may be produced by temperature, partial pressure, or vacuum gradients. Direct contact membrane distillation (DCMD), air-gap membrane distillation (AGMD), sweep gas membrane distillation (SGMD), and vacuum membrane distillation (VMD) are different MD configurations that have reached further applications (see Figure 8).
DCMD and VMD are the most investigated MD configurations [88]. DCMD consists of two liquid phases placed in direct contact on both sides of a microporous membrane. Inside the pore, the formation of a stagnant gas phase occurs. The temperature gradient produces a superficial evaporation on one side of the membrane and a superficial condensation on the opposite side of it. Furthermore, diffusive mass transfer occurs due to the partial pressure gradient of evaporated components. Several authors reported a permeate (evaporate) flux between 5 and 30 (L/hm2) for thermal gradients between 20 and 40 °C [89,90,91,92,93,94]. Additionally, VMD has a productivity greater than DCMD because it allows operating at a pressure lower than 5 kPa, and in this way, obtains high evaporate fluxes [93,94,95].
Recently, Memsys Water Technologies GmbH (Schwabmünchen, Germany) combined the advantages of MED with those of MD processes, resulting in V-MEMD, composed of stages that operate under vacuum with an adequate temperature difference to increase flux through the membrane. In V-MEMD, refrigeration flux in the last stage (highest vacuum) is partly used as preheated feed to the system. The feed passes through each stage, evaporating part way in a cross-flow with the vapor that comes from the thermal system [96].

Forward Osmosis (FO)

This emerging technology has been addressed as a sustainable and cost-efficient solution to classical membrane-based separation technologies such as RO and membrane distillation [97].
Water desalination by means of FO processes consists of the osmotic dilution of the draw solution (DS) and freshwater production from the diluted DS [98,99,100]. FO uses an osmotic pressure differential across the membrane, rather than a hydraulic pressure differential (as in RO), as the driving force for water transport through the membrane. Without the requirement for externally applied hydraulic pressure, FO is installed with a simple and inexpensive low-pressure apparatus, which in turn can reduce the capital costs associated with pumping and system construction [101]. Figure 9 shows FO joined to a conventional desalination process. Re-concentrated DS flux allows pure water flows from feed water. The synchronized operation of the two processes is a key parameter in the whole process design so that desalination is simple, robust, and reliable.
The FO membrane performance mainly depends on the DS properties. The ideal DS should provide high osmotic pressure (higher than the feed osmotic pressure), be inexpensive and nontoxic, provide easy recovery of the solute, be stable, and reduce internal concentration polarization [100].
Recently, FO membrane performance towards desalination was improved by modifying both the DS and FO membranes [103,104,105]. In 2012, Modern Water PLC built “Manipulated Osmosis Desalination (MOD)” in Al Najdah (Oman). MOD was the first commercial FO plant producing 200 m3/d [106].

Reverse Electrodialysis (RED) and Shock Electrodialysis (SED)

RED operates in the same way as ED except for the fact that the voltage is applied in reverse about three to four times an hour with an overall water recovery of 97% [107]. The electrode polarity is reversed at regular intervals for minimizing fouling on the membranes. Therefore, the pretreatments and membrane cleaning are minimal. Rejection of 75–90% is achieved, which depends on ion type and valence, electrical potential, and feeding speed [108]. Industrial applications of RED are brackish water desalination and, at pilot scale, seawater. To the extent that this process can be coupled with a renewable energy source, it will be possible to justify its application for higher salinity waters [109]. On the other hand, Tristan et al. [110] surveyed the life cycle assessment of salinity gradient energy capture by reverse electrodialysis (SGE-RED). They quantified (i) the environmental loads per 1.0 kWh generated by a standalone RED unit and then (ii) the environmental burdens related to the energy provision from an up-scaled RED system to a seawater RO desalination plant per 1.0 m3 of desalted water. The RED unit’s assessment results show that SGE-RED is environmentally competitive with other renewable sources such as photovoltaics or wind. The high salinity solution treated with RED has a lower salt concentration and serves as feed solution to the RO unit to reduce the pump work.
SED is a developing technology that purifies water using polarization zones by the concentration of ions in porous media, adjacent to an ion-selective membrane. A SED cell consists of two ion exchange membranes or electrodes between which feedwater flows through a charged porous medium with thin double layers that act as a “leaky membrane” [111]. When a current passes through the SED cell, an ion-depleted zone is formed along an ion-selective element (the cathode). As the applied voltage is increased, ion concentration near this element approaches zero, and the system can reach the classical diffusion-limited current [106]. Actually, SED operates at small scales, so it holds promise as a decentralized, point-of-use desalination system. SED could be incorporated as a predesalination stage in the RO process, increasing water recovery, decreasing energy consumption, and providing an affordable cost. Alkhadra et al. [112] removed from 70 to 99% of ions of artificial seawater (37,685 ppm) using SED.

1.1.5. Emerging Membrane Processes

Two processes based on polymeric porous media are currently in development and deserve to be considered: graphene oxide membranes (GMs) and aquaporin-based biomimetic membranes (ABMs). On the other hand, hybrid and integrated systems (FO-MD, RO-MD, RO-PRO, etc.) are considered emerging.

Graphene Oxide Membrane (GM)

Graphene oxide (GO), among various forms of nanomaterials, provides tremendous opportunities for rational design and tailoring for solar evaporation and film filtration because of its high absorption, porous structure, high chemical stability, hydrophilicity, and excellent anti-fouling properties [113,114]. To date, GO has been used as an absorber in solar desalination and as a filtration film in membrane desalination [115,116,117,118]. In GMs, there is usually a trade-off between salt rejection and water flux. However, this trade-off can be broken using intercalation, changing the deposition method of GO film or utilizing electrostatic interaction between GO and ions. Currently, GMs are being studied on a laboratory scale [119]. Freire and Pacheco [120] determined that the energy consumption of GMs is less than that of a commercial membrane, since they required lower operating pressures and the coefficient of water mass transport was higher, obtaining permeates with total dissolved solids lower than and equal to 500 ppm.

Aquaporin-Based Biomimetic Membrane (ABM)

Aquaporins (AQPs) are pore-forming proteins in biological cells. These are composed of a bundle of six transmembrane α-helices embedded in the cell membrane. The amino and carboxyl ends face the inside of the cell, whereas the halves resemble each other, apparently repeating a pattern of nucleotides [121,122,123]. Under the right conditions, AQP forms a water channel that selectively transports the water molecules across while excluding ionic species or other polar molecules. This novel property makes AQP a perfect model for the formulation of a low-energy water purification system in seawater desalination [124]. Amy et al. [7] reported that ABMs are being developed as ultrahigh permeability (UHP) RO membranes; with impregnation of AQP (or vesicles) into a polymeric matrix, AQP can provide water channeling/gating, leading to controlled water permeability and ion selectivity.
The major obstacle for ABMs is the scaling up for industry applications, since only small-area membranes have been synthesized due to the highly specialized synthesis techniques. The ABM developed by Zhao et al. [125] had good mechanical stability for periods of weeks to months with stable flux and rejection. Additionally, its permeability was ~40% higher than commercial brackish water RO membrane (BW30) and an order of magnitude higher than seawater RO membrane (SW30HR), which clearly demonstrated the great potential of ABMs for desalination application [124]. ABMs have the potential to reduce energy costs for water treatments [126]. However, these membranes are still at the bench scale and more advancements are necessary to improve their chemical resistance and mechanical strength [127].

Hybrid and Integrated Systems

In recent years, hybrid membrane processes have allowed the achievement of better indicators of the desalination process. For example, Ghaffour et al. [128] presents the state of the art of MD hybrids with different separation processes including RO, PRO, FO, MVC, electrocoagulation, ED, MSF, MED, crystallization, and adsorption with a focus on water production and energy efficiency enhancement. Each of these processes has advantages at the cost of more or less severe drawbacks and their association with MD offers improvement opportunities. Kim et al. [129] proposed a novel module design to integrate FO and MD. The two processes are sealed in one module and operated simultaneously, making the system compact and suitable for a wide range of applications. Results indicated that initial draw solution (MD feed) flow rate and concentration are the most important factors for stable operation of the integrated module.

1.1.6. Technological Improvements Based on Energy Recuperation

Arafat [18] defined two categories for brine energy-recovery systems: the use of pressure exchangers for the direct transfer of the brine pressure to the feed flow, and the use of turbines and pumps that transform rejection pressure energy into mechanical power. Currently, RO plants with a production greater than 3000 m3/d incorporate energy recovery devices, e.g., Pelton turbine, turbocharger, and pressure exchanger (PX device). These devices recover energy from the same desalination process.
The Pelton turbine transforms the rejection pressure into kinetic energy (see Figure 10). The pressurized liquid hits a wheel with vanes that is attached to a high-pressure pump motor. The turbocharger is a compact energy recovery unit, in which a pump and turbine are connected inversely and are provided with a single shaft. The turbine of the turbocharger converts hydraulic pressure energy into mechanical energy that can be used by the pump, allowing an increase in fluid pressure. The PX device directly transfers the high pressure of reject brine to seawater, without previously converting it into mechanical rotation energy. The system uses the principle of positive displacement and isobaric chambers. In fact, the energy savings achieved through these three systems can reach 40%, working with high efficiency (up to 97%) [130]. This represents an SEC close to 2.5 kWh/m3 [131], i.e., according to SEC values reported by Chandwankar and Nowak [24], an SEC equal to that of the MED process and 30% less than the MSF process.

2. Materials and Methods

The comparison method applied in this research can be used in any geographical context. In this case, the expert panel method was applied to determine the best option (technical and economic) of a desalination process for the current Chilean context.
Thermal and membrane processes were evaluated, aiming at applying them to a pilot scale plant that delivers 100 m3/d of purified water starting from seawater at 20 °C and salinity of 34,500 ppm. The desalination plant is supposed to be installed in the coastal zone of the Atacama Region (northern Chile) because:
  • There is abundant solar radiation [133,134] to justify an off-grid installation using photovoltaic panels, or a hybrid arrangement with photovoltaic panels plus accumulation of electrical energy in batteries or with partial supply of electrical energy from the network. Osorio-Aravena et al. [135] reported that, in Chile, renewable electricity will mainly come from solar PV and wind energy technologies. Solar PV and wind energy installed capacities across all sectors would increase from 1.1 GW and 0.8 GW in 2015 to 43.6 GW and 24.8 GW by 2050, respectively. As a consequence, the levelized cost of energy will be reduced by about 25%.
  • Alvez et al. [20] reported that, in Chile, large volumes of water are used in water-scarce regions where mining takes place, alongside agriculture and small communities. This situation has driven a debate around policies to increase the use of seawater to satisfy the water demand of the mining industry.
  • Fragkou and Budds [136] argued that, in Chile, desalination serves to disarticulate drinking water from fresh water, with implications for economic growth, social development, and water policy. They show that desalination entails more than providing additional water to alleviate shortages, and rather constitutes a strategy that permits the reorganization of water sources so as to allow new forms of capital accumulation, through both the water industry as well as the major industries that are threatened by scarcity. They argue that this has three important implications: (1) replacing freshwater with desalinated water for human consumption changes the social relations of control over water, by rendering consumers dependent on desalination plants and their risks, (2) this disarticulation serves to liberate fresh water to sustain the same industries that encroached on drinking water sources, and (3) as a supply-led solution, desalination alleviates some of the water shortages that had been attributed to Chile’s water market model, thereby reducing pressure for reform.
  • Spenceley [137] informed that, in northern Chile, technologically advanced desalination plants are built along the coast, and the desalinated water is moved through an accompanying conveyance system—a complex system of pipelines and pumping infrastructure—over long distances. The resulting brine is released back into the sea through a sophisticated dispersion system designed to reduce brine concentrations to ambient levels efficiently and over the shortest distances possible. Herrera-Leon et al. [138] identified that eleven desalination plants at the industrial scale are operating in Chile (until 2018), producing 5868 l/s of desalinated water. Additionally, there are ten desalination projects in different stages of evaluation, which will increase the desalination capacity by 116.5% to reach a total of 12,706 l/s in the coming years.
  • Due to the high energy demand of desalination techniques, there is a great need for alternatives to reduce the salinity from seawater [139]. In addition, secondary ions, such as calcium and magnesium, in the SW cause scale problems in reverse osmosis plants, mining, and others industries, such as cooling systems. These problems cause increased costs and reduce the efficiency of these processes [140].
  • The use of brackish groundwater often brings risks and obligations to an agricultural system. The application of desalinated water for irrigation can promote soil hydrological functions [141]. However, the disposal of RO concentrate from an inland desalination system can be problematic, and its sustainable management is a major environmental challenge that restricts the widespread application of RO for groundwater desalination [142].
Initially, the classification of information and definition of the technologies to be compared were based on a systematic literature review. Subsequently, the comparative analysis was divided into five categories:
(1)
Technical characteristics,
(2)
Scale-up potential,
(3)
Temperature effect,
(4)
Electrical supply options,
(5)
Economic viability.
Each category was discussed through an expert panel method [143,144,145], composed of five professionals with ample experience in desalination processes. The participants were selected on the basis of professional excellence, landmark publications, and significant teaching experience about desalination processes in Chile. The expert panel analyzed the information from the systematic review and a score was agreed upon according to the criteria shown in Table 2 and Table 3.
Table 2 shows the four criteria levels applied to technical characteristics (category 1): technological development level, operation mode, feasibility of operation with Non-conventional renewable energies (NCRE), pretreatment level, and ease of industrial scaling. The difference in score (1-4 and 1-3) is due to the fact that the expert panel decided to give higher relevance to “technological development level” and “operation mode”. The criteria mean:
  • Incipient: technological development at theoretical and/or laboratory scale.
  • Emerging: technological development prototype at pilot scale.
  • Medium–high: technological development as commercial-scale equipment.
  • High: consolidated technological development and conventional technology, in continuous improvement.
  • Complex: unstable operation.
  • Moderately complex: stable operation.
  • Relatively easy: complicated operation and operation with automation.
  • Easy: very stable process operation and of easy automation.
  • low: the current context does not allow operation with NCRE.
  • medium: the future context does not allow operation with NCRE.
  • high: the current context allows operation with NCRE.
  • demanding: the pretreatment is very necessary to take care of the principal process.
  • moderate: a complex pretreatment is necessary to achieve the goal of the principal process.
  • Simple: a simple pretreatment is necessary to achieve the goal of the principal process.
  • low*: the current context does not allow industrial scaling.
  • medium*: the future context does not allow industrial scaling.
  • high*: the current context allows industrial scaling.
The results of this first evaluation allowed us to determine the technologies with the best characteristics to be applied in Chile and thus focus the analysis of the following categories on those.
Four alternatives were analyzed in categories 2 and 3, based on the results of category 1:
  • Alternative 1 (A1): Nanofiltration;
  • Alternative 2 (A2): Nanomembranes;
  • Alternative 3 (A3): Forward Osmosis + Reverse Osmosis;
  • Alternative 4 (A4): Solar Distillation.
In this research, a comparison between the different alternatives was made considering a production level of 100 m3/d of permeate, with continuous operation for 24 h per day. Table 3 shows the parameters and criteria for the comparative analysis of the industrial scale-up potential (category 2) for the four alternatives. The criteria mean:
  • Low quality: salt content reduced by 50%.
  • Medium quality: salt content reduced by 51–75%.
  • High quality: drinking water quality.
  • Worst: productivity, quality, and cost are lower than RO.
  • Equal: productivity, quality, and cost are similar to RO.
  • Best: productivity, quality, and cost are higher than RO.
  • Worst*: innovation is not better than the current condition.
  • Equal*: innovation equals current condition.
  • Best*: innovation improves the current condition.
  • None: the integration of renewable energy into the desalination process is not possible.
  • Partial: the integration of renewable energy into the desalination process is partially possible.
  • Yes: the integration of renewable energy into the desalination process is possible.
For water quality according to regulations, chloride content is a critical parameter and, if water quality produced is lower than expected, then additional treatment is necessary. The innovations in the operation are the introduction of a new technology to satisfy some need of the desalination process, with science being knowledge, and technology its practice, e.g., seawater feed at a higher temperature.
The temperature effect (category 3) was analyzed for thermal and membrane desalination. Membrane desalination was analyzed based on permeate flux density considering the design equations based on the well-known solution diffusion model.
The alternative that obtained the best results in categories 1, 2, and 3 was analyzed in category 4 (electrical supply options). The analyzed options were:
  • (O1) photovoltaic solar plant + electrical network,
  • (O2) photovoltaic solar plant + batteries,
  • (O3) wind turbine + electrical network.
These options were established to ensure an integrated operational system, constant electricity supply, and utilization of NCRE. The electrical supply options were analyzed based on costs associated with a membrane desalination plant that produces 100 m3/d of permeate and consumes 2.09 kWh/m3 (from pretreatment to posttreatment). This plant must have a daily consumption of 188 kWh, i.e., 7.8 kW/h during the 24 h. The results of the present value of cost (PVC)/m3 of water were analyzed considering an internal return rate of 8% and a 20-year horizon.
Finally, analysis of category 5 (economic viability) applies to the best result obtained in category 4. The economic viability (Capital expenditure (CAPEX) and Operating expenses (OPEX)) of the electrical supply option with the lower associated costs was analyzed. We considered a 20-year time horizon. Costs associated with adduction systems, water distribution to customers, and discharge of rejection into the sea were not considered. CAPEX was determined based on the quotations provided by current suppliers in the market.

3. Results and Discussions

3.1. Comparison of Technologies (Category 1)

Table 4 and Table 5 show a summary of the technical and economic characteristics of desalination processes. This information arose from a systematic literature review. The data indicate that the emerging technology SD has an investment range equal or less than NF, negligible electricity consumption, and a low negative environmental effect. The SD shows the disadvantage of having a very high area-to-volume ratio. Currently, the active layer of NMs has been modified, thus increasing their productivity and selectivity [146,147]. They offer the best technical and economic indicators. The incorporation of aquaporins as an emerging process is the emerging lower-cost technology with the greatest potential for seawater desalination.
Nowadays, MVC, MED, and MSF contribute 40% of desalinated water worldwide (39 Mton/d) [1]. Current trends in RO are energy recovery from concentrated streams and a better membrane permeability, whereas for NF, studies aim at improving the separation efficiency to keep the energy consumption low [59]. On the other hand, FO operates in close cycles with other desalination technologies, thus reducing energy consumption. Finally, GO and ABMs are in the early phases of research, with the expectation that they will achieve their technological maturity and therefore become commercially viable over a 10-year horizon.
The maintenance cost is similar for all technologies, while the technologies that use lower streams presented lower investment and operating costs. This behavior is in agreement with that reported in the literature [88,100].
All desalination technologies have the drawback of returning a concentrated solution to the sea; therefore, the minimization of the effect of a local sea salinity increase due to the reject flow is now the technological challenge.
Table 6 shows the evaluation of the technical characteristics performed by the expert panel. The technologies inserted in the market comply with more than 70% of the technological characteristics. The data suggest that NF and NM have the highest score (after RO), i.e., these technologies have favorable characteristics to stay in the membrane market. This behavior agrees with that reported in the literature [49,85]. RO does not have maximum score due to its demanding level of pretreatment.
Based on the previous results, the technologies chosen by the expert panel to be studied in detail were: NF, NM, FO + RO, and SD. This is because: (1) the technologies supplied with fossil fuel as an energy source were discarded, given the energy context in Northern Chile, (2) SD is an emerging technology with great potential for development in Northern Chile due to greater solar radiation in the area, (3) RO, NF, and NM were the membrane processes that obtained the highest scores, and (4) RO + FO is an emerging process that allows reducing the SEC of traditional RO.

3.2. Industrial Scale-Up Potential (Category 2)

Table 7 shows the evaluation criteria to define the industrial scale-up potential of the four alternatives: NF, NM, FO+RO, and SD. NM achieved the highest score. This evaluation considers that NF increases its productivity by increasing water temperature, but without modifying the saline content of the product. This forces a reprocessing of the permeate at several stages, increasing production cost. Instead, SD achieves high-quality water and productivity, but costs are higher than RO. Finally, NM presents good prospects for industrial scale-up. This technology uses electricity as an energy source; therefore, it would be possible to incorporate renewable energy into its operation. NM is competitive against traditional RO. This behavior agrees with that reported in the literature [146,147,148].

3.3. Temperature Effect on Membrane Productivity (Category 3)

Temperature is a variable that influences desalination processes. The SD process, based on phase change, will increase its productivity (evaporation flow) by less than 5% if the operating temperature increases from 20 to 30 °C. In contrast, NM and RO processes, based on the mass transport through semi-permeable dense polymers, will increase their productivity by around 20%, consistent with what the solution diffusion model predicts [149,150].

3.4. Electrical Supply Options (Category 4)

The preceding results indicate that the best alternative is NM. O1 and O2 consider the use of solar panels installed near the plant. The photovoltaic solar plant loses 20% of its energy, therefore, 9.8 kWh for 24 h are required. These energy requirements are met all year round for 7 h (10 a.m.–5 p.m.) due to solar radiation in Northern Chile. Instead, O3 considers a 25 kW wind turbine. This equipment delivers the energy requirement to the plant for 5 to 7 h (September to April). The conversion efficiency from wind energy to electrical energy is 25–45%.
Table 8 shows the CAPEX calculated for each option (O1, O2, and O3), where O1 presents the lowest CAPEX (equipment).
These results suggest that O1 is the most favorable in the coastal zone of Northern Chile, due to the great potential of solar energy and low investment and maintenance costs.

3.5. Economic Viability (Category 5)

All three options have the seawater desalination system in common, therefore the options share the same economic analysis. Table 9 shows the CAPEX of the seawater desalination plant installation and operation.
For the comparative analysis of the three options, the PVC was used, considering: internal rate of return (IRR) equal to 8%, total evaluation period of 20 years, and operating costs, such as: labor, chemicals, clean-in-place (CIP) cleaning, membrane replacement, spare parts, and plant maintenance, resulting in 0.39 USD/m3 of purified water.
The relevant difference between the options is the cost associated with the type of electrical energy. In this way, the PVC was calculated considering exclusively the item energy, both in CAPEX and OPEX. This implies that O1, O2, and O3 have a PVC of USD 0.085/m3, USD 0.15/m3, and USD 0.115/m3, respectively. In addition, if the plant operates only with an electrical network, then its PVC is USD 0.42/m3.
The results indicate that the use of NM together with a mixed energy supply system (Alternative 2 + Option 1) is profitable. The cost of a seawater desalination system with a daily production of 100 m3/d, and a 40 ft container with thermal and acoustic insulation, is 71–74% of the overall cost of the project. When incorporating the cost of a seawater desalination system with a daily production of 100 m3/d and total operating costs, in this case, a PVC of USD 0.81/m3 is obtained, that is, between 32% and 35% lower compared to the conventional reverse osmosis system, for a plant with a capacity of 100 m3/d.
These results reinforce similar studies by some authors in other geographic locations. For example, Mollahosseini et al. [151] analyzed Iran’s general water background and its renewable energy status, in addition to the potential in renewable energy-assisted desalination (RED). The research suggests that Iran’s potential in RED water production is more than 28 billion m3 in the case where only wind and solar potentials are put into practice. Thus, Iran becomes a prototype in the solutions for water scarcity in cases of proper investment and planning. Jimenez [152] evaluated the feasibility of implementing a desalination plant powered by photovoltaic solar energy in the Colombian Guajira Region. This author determined that the use of renewable energy coupled to the reverse osmosis system is the process that best adjusts to the climatic conditions of the area. Finally, Villagran [153] studied the technical–economic feasibility of the reverse osmosis process with the support of renewable wind and solar photovoltaic energy for a town in Northern Chile. This author determined investments for said plant proportional to an SEC of 2.77, 2.89, and 3.06 kWh/m3.

4. Future Research

In this research, an innovative operational arrangement was identified within the different seawater desalination technologies that can compete in regard to costs (industrial scale) with conventional RO, through a feasibility evaluation of a pilot plant located in Northern Chile. Therefore, the future implementation of the proposed pilot plant will allow obtaining the experimental data.

5. Conclusions

Currently, membrane desalination processes that use energy recuperators have an SEC 50% lower than thermal desalination processes. NM technology exhibits the most favorable technical characteristics and the best economic indexes to consolidate in the desalination market. Even further, NM represents a competitive alternative versus conventional RO process. A profitable desalination process must consider NM, use of energy recuperators, and mixed energy supply. NM + photovoltaic solar plant + electrical network is the most favorable option in the coastal zone of Northern Chile, due to the great potential of solar energy and low investment and maintenance costs. The incorporation of aquaporins as an emerging process is the emerging lower-cost technology with the greatest potential for seawater desalination. A challenge for all desalination technologies is to minimize the effect of local sea salinity increases due to reject streams. This document presents an up-to-date overview of the impact of emerging technologies on desalinated water quality, process costs, productivity, renewable energy use, and separation efficiency. Said information shows that new desalination technologies are more efficient and competitive. NMs are probably the technology that will be used in the future for desalination processes.

Author Contributions

Conceptualization, A.S. and H.V.; methodology, A.S. and O.A.; validation, A.S., O.A., and A.M.; formal analysis, A.S. and H.V.; investigation, A.S. and O.A.; resources, O.A.; data curation, A.S.; writing—original draft preparation, H.V.; writing—review and editing, A.S. and A.M.; visualization, H.V.; supervision, A.S.; project administration, A.S.; funding acquisition, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “Plan de Inversiones Estratégicas en Recursos Hídricos para la Competitividad Meso-Región Norte Chico” development program CORFO-ASOEX, grant number CORFO 16 IFI 65907, Recursos hídricos, 2019. The APC was funded by Vicerrectoría de Investigación, Desarrollo e Innovación (VRIDEI), Universidad de Santiago de Chile and Facultad de Ciencias de la Ingeniería, Universidad Católica del Maule.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data will be available from the corresponding author on reasonable request.

Acknowledgments

S.A., M.A., and H.V. would like to give thanks to DICYT Project 092011SF Universidad de Santiago de Chile for supporting their research in desalination technologies. Valdés, H. would like to give thanks to the fund for attending international congresses from VRIP—Universidad Católica del Maule, 4th International Conference on Desalination using membrane technology, Perth, Australia, 2019.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. International Desalination Association. IDA Water Security Handbook 2019–2020 and GWI DesalData; Media Analytics Ltd.: Oxford, UK, 2019; ISBN 978-1-907467-57-8. [Google Scholar]
  2. Virgili, F.; Pankratz, T.; Gasson, J. IDA Desalination Yearbook 2015–2016; Media Analytics Ltd.: Oxford, UK, 2016; ISBN 1907467408. [Google Scholar]
  3. Wade, N. Distillation plant development and cost update. Desalination 2001, 136, 3–12. [Google Scholar] [CrossRef]
  4. Fajardo-Cadena, A. Desalinización de Agua: ¿una Alternativa Sostenible para la Potabilización del Agua? Bachelor’s Thesis, Fundación Universidad de América, Bogotá, Colombia, 2018. Available online: http://hdl.handle.net/20.500.11839/7141 (accessed on 23 January 2020).
  5. Krishna, H.J. Introduction to Desalination Technologies. Texas Water Development. 2004; Volume 2. Available online: http://www.Twdb.texas.gov/publications/reports/numbered_reports/doc/r363/c1.pdf (accessed on 25 January 2020).
  6. Silva-Pinto, F.; Cunha-Marques, R. Desalination projects economic feasibility: A standardization of cost determinants. Renew. Sustain. Energy Rev. 2017, 78, 904–915. [Google Scholar] [CrossRef]
  7. Amy, G.; Ghaffour, N.; Li, Z.; Francis, L.; Linares, R.V.; Missimer, T.; Lattemann, S. Membrane-based seawater desalination: Present and future prospects. Desalination 2017, 40, 16–21. [Google Scholar] [CrossRef]
  8. Shenvi, S.S.; Isloor, A.M.; Ismail, A.F. A review on RO membrane technology: Developments and challenges. Desalination 2015, 368, 10–26. [Google Scholar] [CrossRef]
  9. Lee, K.; Arnot, T.C.; Mattia, D. A review of reverse osmosis membrane materials for desalination—development to date and future potential. J. Membrane Sci. 2011, 370, 1–22. [Google Scholar] [CrossRef] [Green Version]
  10. Topaloglu, D.; Tilki, Y.M.; Aksu, S.; Yilmaz, T.N.; Celebi, E.E.; Oncel, S.; Aydiner, C. Novel technological solutions for eco-protective water supply by economical and sustainable seawater desalination. Chem. Eng. Res. Des. 2018, 136, 177–198. [Google Scholar] [CrossRef]
  11. Wan, C.; Chung, T. Techno-economic evaluation of various RO+ PRO and RO+ FO integrated processes. Appl. Energ. 2018, 212, 1038–1050. [Google Scholar] [CrossRef]
  12. Ruiz, A.; Andrés, J.; Fernández, J.; Zaragoza, G. Modeling and optimization of a commercial permeate gap spiral wound membrane distillation module for seawater desalination. Desalination 2017, 419, 160–168. [Google Scholar] [CrossRef]
  13. Feitelson, E. On the implications of seawater desalination: Some insights from the Israeli case. In Tapping the Oceans; Edward Elgar Publishing: Cheltenham, UK, 2018. [Google Scholar] [CrossRef]
  14. Sharon, H.; Reddy, K. A review of solar energy driven desalination technologies. Renew. Sustain. Energy Rev. 2015, 41, 1080–1118. [Google Scholar] [CrossRef]
  15. Ghaffour, N.; Bundschuh, J.; Mahmoudi, H.; Goosen, M.F. Renewable energy-driven desalination technologies: A comprehensive review on challenges and potential applications of integrated systems. Desalination 2015, 356, 94–114. [Google Scholar] [CrossRef] [Green Version]
  16. Ghaffour, N.; Missimer, T.; Amy, G. Technical review and evaluation of the economics of water desalination: Current and future challenges for better water supply sustainability. Desalination 2013, 309, 197–207. [Google Scholar] [CrossRef] [Green Version]
  17. Kim, D.; Kim, J.; Shon, H.; Hong, S. Pressure retarded osmosis (PRO) for integrating seawater desalination and wastewater reclamation: Energy consumption and fouling. J. Membrane Sci. 2015, 483, 34–41. [Google Scholar] [CrossRef]
  18. Arafat, H. (Ed.) Desalination Sustainability: A Technical, Socioeconomic, and Environmental Approach; Elsevier: Amsterdam, The Netherlands, 2017; ISBN 9780128097915. [Google Scholar]
  19. Proskynitopoulou, V.; Katsoyiannis, I. Review of recent desalination developments for more efficient drinking water production across the world. New Mater. Compd. Appl. 2018, 2, 179–195. [Google Scholar]
  20. Alvez, A.; Aitken, D.; Rivera, D.; Vergara, M.; McIntyre, N.; Concha, F. At the crossroads: Can desalination be a suitable public policy solution to address water scarcity in Chile’s mining zones? J. Environ. Manag. 2020, 258, 110039. [Google Scholar] [CrossRef] [PubMed]
  21. Gude, V. Desalination and water reuse to address global water scarcity. Rev. Environ. Sci. Bio/Technol. 2017, 16, 591–609. [Google Scholar] [CrossRef]
  22. Day, G.; Schoemaker, P.; Gunther, R. Wharton on Managing Emerging Technologies; John Wiley & Sons: New York, NY, USA, 2000; 460p, ISBN 978-0-471-68939-3. [Google Scholar]
  23. Castillo-Téllez, B.; Pilatowsky-Figueroa, I.; Castillo-Téllez, M.; Lopez-Vidaña, E. Análisis del potencial de la tecnología de congelación/fusión para la desalinización del agua de mar utilizando la energía termo solar. Ingeniantes 2014, 1, 1–11. [Google Scholar]
  24. Chandwankar, R.; Nowak, J. Thermal Processes for Seawater Desalination: Multi-Effect Distillation, Thermal Vapor Compression, Mechanical Vapor Compression, and Multistage Flash; Handbook of Water and Used Water Purification; Springer: Cham, Switzerland, 2019; pp. 1–38. [Google Scholar] [CrossRef]
  25. Alsadaie, S. Design and Operation of Multistage Flash (MSF) Desalination: Advanced Control Strategies and Impact of Fouling. Design Operation and Control of Multistage Flash Desalination Processes: Dynamic Modelling of Fouling, Effect of Non-Condensable Gases on Venting System Design and Implementation of GMC and Fuzzy Control. Ph.D. Thesis, University of Bradford, Bradford, UK, 2017. Available online: http://hdl.handle.net/10454/15924 (accessed on 27 July 2020).
  26. Al-Sofi, M. Fouling phenomena in multi stage flash (MSF) distillers. Desalination 1999, 126, 61–76. [Google Scholar] [CrossRef]
  27. Choi, S. On the brine re-utilization of a multi-stage flashing (MSF) desalination plant. Desalination 2016, 398, 64–76. [Google Scholar] [CrossRef]
  28. Alsadaie, S.; Mujtaba, I. Dynamic modelling of Heat Exchanger fouling in multistage flash (MSF) desalination. Desalination 2017, 409, 47–65. [Google Scholar] [CrossRef] [Green Version]
  29. Voutchkov, N. Desalination Engineering: Planning and Design; McGraw Hill Professional: New York, NY, USA, 2013; ISBN 978-0-07-177716-2. [Google Scholar]
  30. Ghalavand, Y.; Hatamipour, M.S.; Rahimi, A. A review on energy consumption of desalination processes. Desal. Water Treat. 2015, 54, 1526–1541. [Google Scholar] [CrossRef]
  31. Blagin, E.; Biryuk, V.; Anisimov, M.; Shimanov, A.; Gorshkalev, A. Parametric optimization of the MVC desalination plant with thermomechanical compressor. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2018; Volume 339, p. 012032. [Google Scholar] [CrossRef] [Green Version]
  32. Elsayed, M.; Mesalhy, O.; Mohammed, R.; Chow, L. Performance modeling of MED-MVC systems: Exergy-economic analysis. Energy 2019, 166, 552–568. [Google Scholar] [CrossRef]
  33. Ophir, A.; Gendel, A. Steam driven large multi effect MVC (SD MVC) desalination process for lower energy consumption and desalination costs. Desalination 2007, 205, 224–230. [Google Scholar] [CrossRef]
  34. Amiralipour, M.; Kouhikamali, R. Potential analysis and technical-economic optimization of conversion of steam power plant into combined water and power. Appl. Therm. Eng. 2019, 151, 191–198. [Google Scholar] [CrossRef]
  35. Khan, S.; Khan, S.; Haider, S.; El-Leathy, A.; Rana, U.; Danish, S.; Ullah, R. Development and techno-economic analysis of small modular nuclear reactor and desalination system across Middle East and North Africa region. Desalination 2017, 406, 51–59. [Google Scholar] [CrossRef]
  36. Abdallah, A. Modern technology and economical developments in desalination on emphasis of nuclear methodology. Acta Tech. Corvininesis-Bull. Eng. 2018, 11, 49–53. Available online: http://acta.fih.upt.ro/pdf/2018-2/ACTA-2018-2-08.pdf (accessed on 3 August 2020).
  37. Mangal, M.N.; Salinas-Rodriguez, S.G.; Dusseldorp, J.; Kemperman, A.J.; Schippers, J.C.; Kennedy, M.D.; van der Meer, W.G. Effectiveness of antiscalants in preventing calcium phosphate scaling in reverse osmosis applications. J. Membrane Sci. 2021, 119090. [Google Scholar] [CrossRef]
  38. Landsman, M.R.; Lawler, D.F.; Katz, L.E. Application of electrodialysis pretreatment to enhance boron removal and reduce fouling during desalination by nanofiltration/reverse osmosis. Desalination 2020, 491, 114563. [Google Scholar] [CrossRef]
  39. Dhakal, N.; Salinas-Rodriguez, S.G.; Ampah, J.; Schippers, J.C.; Kennedy, M.D. Measuring Biofouling Potential in SWRO Plants with a Flow-Cytometry-Based Bacterial Growth Potential Method. Membranes 2021, 11, 76. [Google Scholar] [CrossRef]
  40. Sperle, P.; Wurzbacher, C.; Drewes, J.E.; Skibinski, B. Reducing the Impacts of Biofouling in RO Membrane Systems through In Situ Low Fluence Irradiation Employing UVC-LEDs. Membranes 2020, 10, 415. [Google Scholar] [CrossRef] [PubMed]
  41. Tai, Z.S.; Abd Aziz, M.H.; Othman, M.H.D.; Mohamed Dzahir, M.I.H.; Hashim, N.A.; Koo, K.N.; Hubadillah, S.K.; Ismail, A.F.; Rahman, M.A.; Jaafar, J. Ceramic membrane distillation for desalination. Sep. Purif. Rev. 2020, 49, 317–356. [Google Scholar] [CrossRef]
  42. Bandar, K.B.; Alsubei, M.D.; Aljlil, S.A.; Darwish, N.B.; Hilal, N. Membrane distillation process application using a novel ceramic membrane for Brackish water desalination. Desalination 2021, 500, 114906. [Google Scholar] [CrossRef]
  43. Voutchkov, N. Energy use for membrane seawater desalination-current status and trends. Desalination 2018, 431, 2–14. [Google Scholar] [CrossRef]
  44. Karabelas, A.; Koutsou, C.; Kostoglou, M.; Sioutopoulos, D. Analysis of specific energy consumption in reverse osmosis desalination processes. Desalination 2018, 431, 15–21. [Google Scholar] [CrossRef]
  45. Rustum, R.; Kurichiyanil, A.M.J.; Forrest, S.; Sommariva, C.; Adeloye, A.J.; Zounemat-Kermani, M.; Scholz, M. Sustainability Ranking of Desalination Plants Using Mamdani Fuzzy Logic Inference Systems. Sustainability 2020, 12, 631. [Google Scholar] [CrossRef] [Green Version]
  46. Karabelas, A.; Koutsou, C.; Sioutopoulos, D.; Plakas, K.; Kostoglou, M. Desalination by Reverse Osmosis. In Sustainable Membrane Technology for Water and Wastewater Treatment; Springer: Singapore, 2017; pp. 155–199. [Google Scholar] [CrossRef]
  47. Mazlan, N.; Peshev, D.; Livingston, A. Energy consumption for desalination—A comparison of forward osmosis with reverse osmosis, and the potential for perfect membranes. Desalination 2016, 377, 138–151. [Google Scholar] [CrossRef] [Green Version]
  48. Mohammad, A.; Teow, Y.; Ang, W.; Chung, Y.; Oatley-Radcliffe, D.; Hilal, N. Nanofiltration membranes review: Recent advances and future prospects. Desalination 2015, 356, 226–254. [Google Scholar] [CrossRef]
  49. Eke, J.; Yusuf, A.; Giwa, A.; Sodiq, A. The global status of desalination: An assessment of current desalination technologies, plants and capacity. Desalination 2020, 495, 114633. [Google Scholar] [CrossRef]
  50. Wright, N.; Shah, S.; Amrose, S. A robust model of brackish water electrodialysis desalination with experimental comparison at different size scales. Desalination 2018, 443, 27–43. [Google Scholar] [CrossRef]
  51. Chua, S.C.; Isa, M.H.; Ho, Y.C. Electrodialysis (ED): A review on the fundamental concept, advantages, limitations and future trend. Platform A J. Sci. Technol. 2020, 3, 14–22. Available online: http://myjms.mohe.gov.my/index.php/pjst/article/view/6146 (accessed on 27 January 2021).
  52. Gurreri, L.; Tamburini, A.; Cipollina, A.; Micale, G. Electrodialysis Applications in Wastewater Treatment for Environmental Protection and Resources Recovery: A Systematic Review on Progress and Perspectives. Membranes 2020, 10, 146. [Google Scholar] [CrossRef]
  53. Ismail, A.; Matsuura, T. Progress in transport theory and characterization method of Reverse Osmosis (RO) membrane in past fifty years. Desalination 2018, 434, 2–11. [Google Scholar] [CrossRef]
  54. Dennison, J.; Xie, X.; Murphy, C.; Cahill, D. Density, Elastic Constants, and Thermal Conductivity of Interfacially Polymerized Polyamide Films for Reverse Osmosis Membranes. ACS Appl. Nano Mater. 2018, 1, 5008–5018. [Google Scholar] [CrossRef]
  55. Li, Y.; Pan, G.; Wang, J.; Zhang, Y.; Shi, H.; Yu, H.; Liu, Y. Tailoring the polyamide active layer of thin-film composite forward osmosis membranes with combined cosolvents during interfacial polymerization. Ind. Eng. Chem. Res. 2020, 59, 8230–8242. [Google Scholar] [CrossRef]
  56. Ruiz-García, A.; Nuez, I. Performance Assessment of SWRO Spiral-Wound Membrane Modules with Different Feed Spacer Dimensions. Processes 2020, 8, 692. [Google Scholar] [CrossRef]
  57. Taherinejad, M.; Derakhshan, S.; Yavarinasab, A. Hydrodynamic analysis of spiral wound reverse osmosis membrane recovery fraction and permeate water flow rate. Desalination 2017, 411, 59–68. [Google Scholar] [CrossRef]
  58. Vrouwenvelder, J.; Kruithof, J.; van Loosdrecht, M. Biofouling of Spiral Wound Membrane Systems; IWA Publishing: London, UK; New York, NY, USA, 2019; ISBN 978178040. [Google Scholar]
  59. Kingsbury, R.S.; Wang, J.; Coronell, O. Comparison of water and salt transport properties of ion exchange, reverse osmosis, and nanofiltration membranes for desalination and energy applications. J. Membrane Sci. 2020, 604, 117998. [Google Scholar] [CrossRef]
  60. Álvarez, J.R.; Antón, F.E.; Álvarez-García, S.; Luque, S. Treatment of Aqueous Effluents from Steel Manufacturing with High Thiocyanate Concentration by Reverse Osmosis. Membranes 2020, 10, 437. [Google Scholar] [CrossRef]
  61. Marecka-Migacz, A.; Mitkowski, P.T.; Nędzarek, A.; Różański, J.; Szaferski, W. Effect of pH on Total Volume Membrane Charge Density in the Nanofiltration of Aqueous Solutions of Nitrate Salts of Heavy Metals. Membranes 2020, 10, 235. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, H.; He, Q.; Luo, J.; Wan, Y.; Darling, S.B. Sharpening nanofiltration: Strategies for enhanced membrane selectivity. ACS Appl. Mater. Inter. 2020, 12, 39948–39966. [Google Scholar] [CrossRef]
  63. Zhou, X.; Wang, Z.; Epsztein, R.; Zhan, C.; Li, W.; Fortner, J.D.; Pham, T.A.; Kim, J.-H.; Elimelech, M. Intrapore energy barriers govern ion transport and selectivity of desalination membranes. Sci. Adv. 2020, 6, eabd9045. [Google Scholar] [CrossRef] [PubMed]
  64. Prajapati, M.; Shah, M.; Soni, B.; Parikh, S.; Sircar, A.; Balchandani, S.; Thakore, S.; Tala, M. Geothermal-solar integrated groundwater desalination system: Current status and future perspective. Ground. Sustain. Dev. 2020, 100506. [Google Scholar] [CrossRef]
  65. Gorjian, S.; Ghobadian, B. Solar desalination: A sustainable solution to water crisis in Iran. Renew. Sustain. Energy Rev. 2015, 48, 571–584. [Google Scholar] [CrossRef]
  66. Pandey, K.; Chaurasiya, R. A review on analysis and development of solar flat plate collector. Renew. Sustain. Energy Rev. 2017, 67, 641–650. [Google Scholar] [CrossRef]
  67. Martinopoulos, G.; Ikonomopoulos, A.; Tsilingiridis, G. Initial evaluation of a phase changes solar collector for desalination applications. Desalination 2016, 399, 165–170. [Google Scholar] [CrossRef]
  68. Gude, V. Energy Storage for Desalination. In Renewable Energy Powered Desalination Handbook; Butterworth-Heinemann: Oxford, UK, 2018; pp. 377–414. [Google Scholar] [CrossRef]
  69. Razmi, A.; Soltani, M.; Tayefeh, M.; Torabi, M.; Dusseault, M. Thermodynamic analysis of compressed air energy storage (CAES) hybridized with a multi-effect desalination (MED) system. Energy Convers. Manag. 2019, 199, 112047. [Google Scholar] [CrossRef]
  70. Panchal, H. Use of thermal energy storage materials for enhancement in distillate output of solar still: A review. Renew. Sustain. Energy Rev. 2016, 61, 86–96. [Google Scholar] [CrossRef]
  71. Omar, A.; Nashed, A.; Li, Q.; Leslie, G.; Taylor, R.A. Pathways for integrated concentrated solar power-Desalination: A critical review. Renew. Sustain. Energy Rev. 2020, 119, 109609. [Google Scholar] [CrossRef]
  72. Palenzuela, P.; Alarcón-Padilla, D.; Zaragoza, G.; Blanco, J. Comparison between CSP + MED and CSP + RO in Mediterranean area and MENA region: Techno-economic analysis. Energy Procedia 2015, 69, 1938–1947. [Google Scholar] [CrossRef] [Green Version]
  73. Li, S.; Cai, Y.; Schäfer, A.; Richards, B. Renewable energy powered membrane technology: A review of the reliability of photovoltaic-powered membrane system components for brackish water desalination. Appl. Energy 2019, 253, 113524. [Google Scholar] [CrossRef]
  74. Alsarayreh, A.; Majdalawi, M.; Bhandari, R. Techno-Economic study of PV Powered Brackish Water Reverse Osmosis Desalination Plant in the Jordan Valley. Int. J. Therm. Environ. Eng. 2017, 14, 83–88. [Google Scholar] [CrossRef]
  75. Mehrabian-Nejad, H.; Farhangi, B.; Farhangi, S. Application of PV and Solar Energy in Water Desalination System. J. Sol. Energy Res. 2017, 2, 13–18. Available online: https://jser.ut.ac.ir/article_62441_50a967809fe6d8e 41937b9a68e828f42.pdf (accessed on 18 September 2020).
  76. Majdi, H. Design and sizing of small–scale photovoltaic (PV) cells powered reverse osmosis (RO) desalination system for water supply in remote locations. Iraqi J. Mech. Mater. Eng. 2016, 16, 350–365. Available online: https://www.iasj.net/iasj?func=fulltext&aId=120423 (accessed on 18 September 2020).
  77. Rezk, H.; Abdelkareem, M.; Ghenai, C. Performance evaluation and optimal design of stand-alone solar PV-battery system for irrigation in isolated regions: A case study in Al Minya (Egypt). Sustain. Energy Technol. 2019, 36, 100556. [Google Scholar] [CrossRef]
  78. Gökçek, M. Integration of hybrid power (wind-photovoltaic-diesel-battery) and seawater reverse osmosis systems for small-scale desalination applications. Desalination 2018, 435, 210–220. [Google Scholar] [CrossRef]
  79. Gude, V. Geothermal source potential for water desalination–Current status and future perspective. Renew. Sustain. Energy Rev. 2016, 57, 1038–1065. [Google Scholar] [CrossRef]
  80. Assad, M.E.H.; AlShabi, M.; Sahlolbei, A.; Hmida, A.; Khuwaileh, B. Geothermal energy use in seawater desalination. Proc. SPIE 2020, 11387, 1138716. [Google Scholar] [CrossRef]
  81. Ghaffour, N.; Lattemann, S.; Missimer, T.; Ng, K.C.; Sinha, S.; Amy, G. Renewable energy-driven innovative energy-efficient desalination technologies. Appl. Energy 2014, 136, 1155–1165. [Google Scholar] [CrossRef] [Green Version]
  82. Zhao, D.; Japip, S.; Zhang, Y.; Weber, M.; Maletzko, C.; Chung, T. Emerging thin-film nanocomposite (TFN) membranes for reverse osmosis: A review. Water Res. 2020, 115557. [Google Scholar] [CrossRef] [PubMed]
  83. Safarpour, M.; Khataee, A.; Vatanpour, V. Thin film nanocomposite reverse osmosis membrane modified by reduced graphene oxide/TiO2 with improved desalination performance. J. Membrane Sci. 2015, 489, 43–54. [Google Scholar] [CrossRef]
  84. Ganesan, J.; Gandhi, M.P.; Nagendran, M.; Li, B.; Nair, V.; Chellam, P.V. Functional Properties of Nanoporous Membranes for the Desalination of Water. Environ. Nanotechnol. 2020, 4, 131–163. [Google Scholar] [CrossRef]
  85. Voutchkov, N. Desalination Project Cost Estimating and Management; Taylor & Francis Group, CRC Press: Boca Raton, FL, USA, 2019; ISBN 978-0-8153-7414-5. [Google Scholar]
  86. Khorshidi, B.; Biswas, I.; Ghosh, T.; Thundat, T.; Sadrzadeh, M. Robust fabrication of Thin Film Polyamide-TiO2 nanocomposite membranes with enhanced thermal sand anti-biofouling propensity. Sci. Rep. 2018, 8, 784. [Google Scholar] [CrossRef]
  87. Anvari, A.; Yancheshme, A.A.; Kekre, K.M.; Ronen, A. State-of-the-art methods for overcoming temperature polarization in membrane distillation process: A review. J. Membrane Sci. 2020, 118413. [Google Scholar] [CrossRef]
  88. Gonzalez, D.; Amigo, J.; Suárez, F. Membrane distillation: Perspectives for sustainable and improved desalination. Renew. Sustain. Energy Rev. 2017, 80, 238–259. [Google Scholar] [CrossRef]
  89. Boubakri, A.; Elgharbi, S.; Bouguecha, S.A.T.; Hafiane, A. Energetic Performance and Permeate Flux Investigation of Direct-Contact Membrane Distillation for Seawater Desalination. Chem. Eng. Technol. 2020, 43, 2457–2468. [Google Scholar] [CrossRef]
  90. Chen, Y.-H.; Hung, H.-G.; Ho, C.-D.; Chang, H. Economic Design of Solar-Driven Membrane Distillation Systems for Desalination. Membranes 2021, 11, 15. [Google Scholar] [CrossRef]
  91. Lisboa, K.M.; de Moraes, D.B.; Naveira-Cotta, C.P.; Cotta, R.M. Analysis of the membrane effects on the energy efficiency of water desalination in a direct contact membrane distillation (DCMD) system with heat recovery. Appl. Therm. Eng. 2021, 182, 116063. [Google Scholar] [CrossRef]
  92. Ameen, N.A.M.; Ibrahim, S.S.; Alsalhy, Q.F.; Figoli, A. Highly Saline Water Desalination Using Direct Contact Membrane Distillation (DCMD): Experimental and Simulation Study. Water 2020, 12, 1575. [Google Scholar] [CrossRef]
  93. Donato, L.; Garofalo, A.; Drioli, E.; Alharbi, O.; Aljlil, S.A.; Criscuoli, A.; Algieri, C. Improved performance of vacuum membrane distillation in desalination with zeolite membranes. Sep. Purif. Technol. 2020, 237, 116376. [Google Scholar] [CrossRef]
  94. Kim, H.; Yun, T.; Hong, S.; Lee, S. Experimental and theoretical investigation of a high performance PTFE membrane for vacuum-membrane distillation. J. Membrane Sci. 2020, 617, 118524. [Google Scholar] [CrossRef]
  95. Cao, W.; Liu, Q.; Wang, Y.; Mujtaba, I. Modeling and simulation of VMD desalination process by ANN. Comput. Chem. Eng. 2016, 84, 96–103. [Google Scholar] [CrossRef] [Green Version]
  96. Memsys. Membrane Distillation Technology. Available online: https://www.memsys.eu/ (accessed on 28 June 2020).
  97. Goh, P.; Ismail, A.; Ng, B.; Abdullah, M. Recent Progresses of Forward Osmosis Membranes Formulation and Design for Wastewater Treatment. Water 2019, 11, 2043. [Google Scholar] [CrossRef] [Green Version]
  98. Suzaimi, N.D.; Goh, P.S.; Ismail, A.F.; Mamah, S.C.; Malek, N.A.N.N.; Lim, J.W.; Wong, K.C.; Hilal, N. Strategies in Forward Osmosis Membrane Substrate Fabrication and Modification: A Review. Membranes 2020, 10, 332. [Google Scholar] [CrossRef]
  99. Mohammadifakhr, M.; de Grooth, J.; Roesink, H.D.W.; Kemperman, A.J.B. Forward Osmosis: A Critical Review. Processes 2020, 8, 404. [Google Scholar] [CrossRef] [Green Version]
  100. Ibrahim, G.; Isloor, A.; Yuliwati, E. A Review: Desalination by Forward Osmosis. In Current Trends and Future Developments on (Bio-)Membranes; Elsevier: Amsterdam, The Netherlands, 2019; pp. 199–214. [Google Scholar] [CrossRef]
  101. Valladares, R.; Li, Z.; Abu-Ghdaib, M.; Wei, C.-H.; Amy, G.; Vrouwenvelder, J. Water harvesting from municipal wastewater via osmotic gradient: An evaluation of process performance. J. Membrane Sci. 2013, 447, 50–56. [Google Scholar] [CrossRef]
  102. Tiraferri, A. Forward osmosis for water treatment and desalination. In Current Trends and Future Developments on (Bio-)Membranes; Elsevier: Amsterdam, The Netherlands, 2020; pp. 53–81. [Google Scholar] [CrossRef]
  103. Johnson, D.; Suwaileh, W.; Mohammed, A.; Hilal, N. Osmotic’s potential: An overview of draw solutes for forward osmosis. Desalination 2018, 434, 100–120. [Google Scholar] [CrossRef] [Green Version]
  104. Ray, S.; Chen, S.; Nguyen, N.; Nguyen, H.; Dan, N.; Thanh, B. Exploration of polyelectrolyte incorporated with Triton-X 114 surfactant based osmotic agent for forward osmosis desalination. J. Environ. Manag. 2018, 209, 346–353. [Google Scholar] [CrossRef]
  105. Cheng, W.; Lu, X.; Yang, Y.; Jiang, J.; Ma, J. Influence of composition and concentration of saline water on cation exchange behavior in forward osmosis desalination. Water Res. 2018, 137, 9–17. [Google Scholar] [CrossRef]
  106. Nicoll, P. Forward osmosis as a pre-treatment to reverse osmosis. In The International Desalination Association World Congress on Desalination and Water Reuse; International Desalination Association: Tianjin, China, 2013. [Google Scholar]
  107. Xu, P.; Cath, T.; Robertson, A.; Reinhard, M.; Leckie, J.; Drewes, J. Critical review of desalination concentrate management, treatment and beneficial use. Environ. Eng. Sci. 2013, 30, 502–514. [Google Scholar] [CrossRef] [Green Version]
  108. Hanrahan, C.; Karimi, L.; Ghassemi, A.; Sharbat, A. High-recovery electrodialysis reversal for the desalination of inland brackish waters. Desalin. Water Treat. 2016, 57, 11029–11039. [Google Scholar] [CrossRef]
  109. Zoungrana, A.; Çakmakci, M. From non-renewable energy to renewable by harvesting salinity gradient power by reverse electrodialysis: A review. Int. J. Energy Res. 2020. [Google Scholar] [CrossRef]
  110. Tristán, C.; Rumayor, M.; Dominguez-Ramos, A.; Fallanza, M.; Ibáñez, R.; Ortiz, I. Life cycle assessment of salinity gradient energy recovery by reverse electrodialysis in a seawater reverse osmosis desalination plant. Sustain. Energy Fuels 2020, 4, 4273–4284. [Google Scholar] [CrossRef]
  111. Deng, D.; Aouad, W.; Braff, W.; Schlumpberger, S.; Suss, M.; Bazant, M. Water purification by shock electrodialysis: Deionization, filtration, separation, and disinfection. Desalination 2015, 357, 77–83. [Google Scholar] [CrossRef] [Green Version]
  112. Alkhadra, M.; Gao, T.; Conforti, K.; Tian, H.; Bazant, M. Small-scale desalination of seawater by shock electrodialysis. Desalination 2020, 476, 114219. [Google Scholar] [CrossRef]
  113. Johnson, D.J.; Hilal, N. Can graphene and graphene oxide materials revolutionize desalination processes? Desalination 2020, 114852. [Google Scholar] [CrossRef]
  114. Homaeigohar, S.; Elbahri, M. Graphene membranes for water desalination. NPG Asia Mater. 2017, 9, e427. [Google Scholar] [CrossRef] [Green Version]
  115. Mao, Y.; Huang, Q.; Meng, B.; Zhou, K.; Liu, G.; Gugliuzza, A.; Drioli, E.; Jin, W. Roughness-enhanced hydrophobic graphene oxide membrane for water desalination via membrane distillation. J. Membrane Sci. 2020, 611, 118364. [Google Scholar] [CrossRef]
  116. Thakur, A.K.; Sathyamurthy, R.; Sharshir, S.W.; Ahmed, M.S.; Hwang, J.Y. A novel reduced graphene oxide based absorber for augmenting the water yield and thermal performance of solar desalination unit. Mater. Lett. 2020, 286, 128867. [Google Scholar] [CrossRef]
  117. Yang, Y.; Zhao, R.; Zhang, T.; Zhao, K.; Xiao, P.; Ma, Y.; Ajayan, P.; Shi, G.; Chen, Y. Graphene-based standalone solar energy converter for water desalination and purification. ACS Nano 2018, 12, 829–835. [Google Scholar] [CrossRef] [PubMed]
  118. Hong, S.; Constans, C.; Surmani-Martins, M.; Seow, Y.; Guevara-Carrió, J.; Garaj, S. Scalable graphene-based membranes for ionic sieving with ultrahigh charge selectivity. Nano Lett. 2017, 17, 728–732. [Google Scholar] [CrossRef] [PubMed]
  119. Li, X.; Zhu, B.; Zhu, J. Graphene oxide based materials for desalination. Carbon 2019, 146, 320–328. [Google Scholar] [CrossRef]
  120. Freire, T.; Pacheco, C. Estudio del Consumo de Energía en el Proceso de Osmosis Inversa Utilizando un Filtro de Membrana de Grafeno para la Desalinización del Agua del mar. Bachelor’s Thesis, UCE, Boca Raton, FL, USA, 2017; 125p. Available online: http://www.dspace.uce.edu.ec/handle/25000/12918 (accessed on 7 October 2020).
  121. Porter, C.J.; Werber, J.R.; Zhong, M.; Wilson, C.J.; Elimelech, M. Pathways and Challenges for Biomimetic Desalination Membranes with Sub-Nanometer Channels. ACS Nano 2020, 14, 10894–10916. [Google Scholar] [CrossRef]
  122. Fu, D.; Lu, M. The structural basis of water permeation and proton exclusion in aquaporins. Mol. Membr. Biol. 2007, 24, 366–374. [Google Scholar] [CrossRef] [Green Version]
  123. Gonen, T.; Walz, T. The structure of aquaporins. Q. Rev. Biophys. 2006, 39, 361–396. [Google Scholar] [CrossRef]
  124. Teow, Y.; Mohammad, A. New generation nanomaterials for water desalination: A review. Desalination 2019, 451, 2–17. [Google Scholar] [CrossRef]
  125. Zhao, Y.; Qiu, C.; Li, X.; Vararattanavech, A.; Shen, W.; Torres, J.; Hélix-Nielsen, C.; Wang, R.; Hu, X.; Fane, A.; et al. Synthesis of robust and high-performance aquaporin-based biomimetic membranes by interfacial polymerization-membrane preparation and RO performance characterization. J. Membrane Sci. 2012, 423, 422–428. [Google Scholar] [CrossRef]
  126. Wang, R.; Fane, A. Aquaporins (AQP)-Based Biomimetic Membranes for Water Reuse and Desalination in “Advanced Membrane Technology VII”. ECI Symposium Series. 2016. Available online: https://dc.engconfintl.org/membrane_technology_vii/40 (accessed on 7 October 2020).
  127. Gude, V.; Fthenakis, V. Energy efficiency and renewable energy utilization in desalination systems. Prog. Energy 2020. [Google Scholar] [CrossRef]
  128. Ghaffour, N.; Soukane, S.; Lee, J.G.; Kim, Y.; Alpatova, A. Membrane distillation hybrids for water production and energy efficiency enhancement: A critical review. Appl. Energy 2019, 254, 113698. [Google Scholar] [CrossRef]
  129. Kim, Y.; Li, S.; Francis, L.; Li, Z.; Linares, R.V.; Alsaadi, A.S.; Abu-Ghdaib, M.; Son, H.S.; Amy, G.; Ghaffour, N. Osmotically and thermally isolated forward osmosis–membrane distillation (FO–MD) integrated module. Environ. Sci. Technol. 2019, 53, 3488–3498. [Google Scholar] [CrossRef]
  130. Urrea, S.; Reyes, F.; Suárez, B.; Juan, A. Technical review, evaluation and efficiency of energy recovery devices installed in the Canary Islands desalination plants. Desalination 2019, 450, 54–63. [Google Scholar] [CrossRef]
  131. Peñate, B.; García-Rodríguez, L. Energy optimization of existing SWRO (seawater reverse osmosis) plants with ERT (energy recovery turbines): Technical and thermoeconomic assessment. Energy 2011, 36, 613–626. [Google Scholar] [CrossRef]
  132. El-Emam, R.S.; Dincer, I. Thermodynamic and thermoeconomic analyses of seawater reverse osmosis desalination plant with energy recovery. Energy 2014, 64, 154–163. [Google Scholar] [CrossRef]
  133. Castillejo-Cuberos, A.; Escobar, R. Understanding solar resource variability: An in-depth analysis, using Chile as a case of study. Renew. Sustain. Energy Rev. 2020, 120, 109664. [Google Scholar] [CrossRef]
  134. Merino, I.; Herrera, I.; Valdés, H. Environmental Assessment of Energy Scenarios for a Low-Carbon Electrical Network in Chile. Sustainability 2019, 11, 5066. [Google Scholar] [CrossRef] [Green Version]
  135. Osorio-Aravena, J.C.; Aghahosseini, A.; Bogdanov, D.; Caldera, U.; Muñoz-Cerón, E.; Breyer, C. Transition toward a fully renewable based energy system in Chile by 2050 across power, heat, transport and desalination sectors. Int. J. Sustain. Energy Plan Manag. 2020, 25, 77–94. [Google Scholar] [CrossRef]
  136. Fragkou, M.C.; Budds, J. Desalination and the disarticulation of water resources: Stabilising the neoliberal model in Chile. T. I. Brit. Geogr. 2020, 45, 448–463. [Google Scholar] [CrossRef]
  137. Spenceley, J. Sustainable Water Supply for Chile’s Copper Mines. Eng. Min. J. 2020, 221, 38–41. Available online: https://search.proquest.com/scholarly-journals/sustainable-water-supply-chiles-copper-mines/docview/2351559852/se-2?accountid=170518 (accessed on 20 January 2021).
  138. Herrera-Leon, S.; Cruz, C.; Kraslawski, A.; Cisternas, L.A. Current situation and major challenges of desalination in Chile. Desalin. Water Treat 2019, 171, 93–104. [Google Scholar] [CrossRef]
  139. Karagiannis, I.C.; Soldatos, P.G. Water desalination cost literature: Review and assessment. Desalination 2008, 223, 448–456. [Google Scholar] [CrossRef]
  140. Arias, D.; Cisternas, L.A.; Rivas, M. Biomineralization of calcium and magnesium crystals from seawater by halotolerant bacteria isolated from Atacama Salar (Chile). Desalination 2017, 405, 1–9. [Google Scholar] [CrossRef]
  141. Assouline, S.; Russo, D.; Silber, A.; Or, D. Balancing water scarcity and quality for sustainable irrigated agriculture. Water Resour. Res. 2015, 51, 3419–3436. [Google Scholar] [CrossRef]
  142. Baath, G.S.; Shukla, M.; Bosland, P.W.; Walker, S.J.; Saini, R.K.; Shaw, R. Water Use and Yield Responses of Chile Pepper Cultivars Irrigated with Brackish Groundwater and Reverse Osmosis Concentrate. Horticulturae 2020, 6, 27. [Google Scholar] [CrossRef]
  143. Adamec, J.; Janoušková, S.; Hák, T. How to Measure Sustainable Housing: A Proposal for an Indicator-Based Assessment Tool. Sustainability 2021, 13, 1152. [Google Scholar] [CrossRef]
  144. Nind, M.; Lewthwaite, S. Methods that teach: Developing pedagogic research methods, developing pedagogy. Int. J. Res. Method Educ. 2018, 41, 398–410. [Google Scholar] [CrossRef] [Green Version]
  145. Valdés, H.; Correa, C.; Mellado, F. Proposed Model of Sustainable Construction Skills for Engineers in Chile. Sustainability 2018, 10, 3093. [Google Scholar] [CrossRef] [Green Version]
  146. Manikandan, S.; Karmegam, N.; Subbaiya, R.; Devi, G.K.; Arulvel, R.; Ravindran, B.; Awasthi, M.K. Emerging nano-structured innovative materials as adsorbents in wastewater treatment. Bioresour. Technol. 2020, 124394. [Google Scholar] [CrossRef] [PubMed]
  147. Uddin, M.N.; Alamir, M.; Muppalla, H.; Rahman, M.M.; Asmatulu, R. Nanomembranes for sustainable fresh water production. In Proceedings of the International Conference on Mechanical, Industrial and Energy Engineering, ICMIEE18-KN01, Khulna, Bangladesh, 23–24 December 2018. [Google Scholar]
  148. Rehman, F.; Thebo, K.H.; Aamir, M.; Akhtar, J. Nanomembranes for water treatment. In Nanotechnology in the Beverage Industry; Elsevier: Amsterdam, The Netherlands, 2020; pp. 207–240. [Google Scholar] [CrossRef]
  149. Jouzdani, S.M.; Zerafat, M.M.; Arasteh, P.D.; Vagharfard, H. Evaluation of environmental indicators of RO seawater desalination: Case study coastal strip of Hormozgan province, Iran. J. Water Supply Res. Technol.—AQUA 2020, 69, 694–703. [Google Scholar] [CrossRef]
  150. Ruiz-García, A.; Nuez, I. Long-term intermittent operation of a full-scale BWRO desalination plant. Desalination 2020, 489, 114526. [Google Scholar] [CrossRef]
  151. Mollahosseini, A.; Abdelrasoul, A.; Sheibany, S.; Amini, M.; Salestan, S. Renewable energy-driven desalination opportunities—A case study. J. Environ. Manag. 2019, 239, 187–197. [Google Scholar] [CrossRef] [PubMed]
  152. Jimenez, H. Evaluación de la Viabilidad para la Implementación de Plantas Desalinizadoras Usando Energías Renovables en la Región de la Guajira Colombiana (Monografía). Fundación Universidad de América. 2019. Available online: http://hdl.handle.net/20.500.11839/7253 (accessed on 19 December 2020).
  153. Villagran, S. Factibilidad de Desalinización de Agua de mar para Pequeñas Comunidades del Norte de Chile. Bachelor’s Thesis, Universidad de Chile, Santiago, Chile, 2017. Available online: http://repositorio.uchile.cl/handle/2250/145387 (accessed on 20 December 2020).
Figure 1. Schematic classification of desalination technologies.
Figure 1. Schematic classification of desalination technologies.
Membranes 11 00180 g001
Figure 2. Conceptual diagram of the conventional thermal desalination technologies (Mentioned references [3,24,27,28,29,30,31,32,33,34,35,36]).
Figure 2. Conceptual diagram of the conventional thermal desalination technologies (Mentioned references [3,24,27,28,29,30,31,32,33,34,35,36]).
Membranes 11 00180 g002
Figure 3. Conceptual diagram of the conventional membrane technologies applied to desalination (Mentioned references [53,54,55,56,57,58,59,60,61,62,63]).
Figure 3. Conceptual diagram of the conventional membrane technologies applied to desalination (Mentioned references [53,54,55,56,57,58,59,60,61,62,63]).
Membranes 11 00180 g003
Figure 4. Conceptual diagram depicting the classification of solar collector technologies.
Figure 4. Conceptual diagram depicting the classification of solar collector technologies.
Membranes 11 00180 g004
Figure 5. Schematic view of (a) Concentrating solar power + Reverse osmosis (CSP/RO) plant, (b) Photovoltaic solar panel + Reverse osmosis (PV/RO) plant.
Figure 5. Schematic view of (a) Concentrating solar power + Reverse osmosis (CSP/RO) plant, (b) Photovoltaic solar panel + Reverse osmosis (PV/RO) plant.
Membranes 11 00180 g005
Figure 6. (a) Combined-cycle geothermal and solar-powered desalination. The system is powered by solar energy during daylight hours and by geothermal energy during nighttime and cloudy days. (b) Principle of direct contact membrane distillation process. Based on Ghaffour et al. [81].
Figure 6. (a) Combined-cycle geothermal and solar-powered desalination. The system is powered by solar energy during daylight hours and by geothermal energy during nighttime and cloudy days. (b) Principle of direct contact membrane distillation process. Based on Ghaffour et al. [81].
Membranes 11 00180 g006
Figure 7. Example of scheme of a cross-flow desalination system for nanomembranes. Based on Safarpour et al. [83].
Figure 7. Example of scheme of a cross-flow desalination system for nanomembranes. Based on Safarpour et al. [83].
Membranes 11 00180 g007
Figure 8. Schematic illustration of different approaches for the conventional MD process: direct contact membrane distillation (DCMD), air-gap membrane distillation (AGMD), sweep gas membrane distillation (SGMD), and vacuum membrane distillation (VMD). Based on Anvari et al. [87].
Figure 8. Schematic illustration of different approaches for the conventional MD process: direct contact membrane distillation (DCMD), air-gap membrane distillation (AGMD), sweep gas membrane distillation (SGMD), and vacuum membrane distillation (VMD). Based on Anvari et al. [87].
Membranes 11 00180 g008
Figure 9. Schematic view of FO pretreatment in hybrid systems for desalination. Based on Tiraferri [102].
Figure 9. Schematic view of FO pretreatment in hybrid systems for desalination. Based on Tiraferri [102].
Membranes 11 00180 g009
Figure 10. Schematic diagram of SWRO desalination plant with energy-recovery turbine. Based on El-Emam & Dincer [132].
Figure 10. Schematic diagram of SWRO desalination plant with energy-recovery turbine. Based on El-Emam & Dincer [132].
Membranes 11 00180 g010
Table 1. Summary of conventional and emerging technologies for desalination according to physicochemical process and external gradient.
Table 1. Summary of conventional and emerging technologies for desalination according to physicochemical process and external gradient.
TechnologyExternal GradientPhysicochemical Process
ConventionalMulti-stage flash distillation (MSF)ThermalPhase change
Multi-effect distillation (MED)
Mechanical vapor compression (MVC)
Freezing
Reverse osmosis (RO)Total pressurePhysical,
Nanofiltration (NF)without phase change
Electrodialysis (ED)ElectricElectric,
without phase change
Ion exchange (IE)ChemicalChemical,
without phase change
EmergingSolar distillation (SD)ThermalPhase change
Solar thermal with RO
Solar photovoltaic with RO
Geothermal desalination
Nanomembranes (NMs)ChemicalPhysical, membrane properties improved
Membrane distillation (MD)ThermalPhase change
Forward osmosis (FO)Osmotic pressureChemical,
with phase change
Reverse electrodialysis (RED)ElectricElectric
Shock electrodialysis (SED)without phase change
Biomimetics (aquaporins)ChemicalChemical
Graphene membrane (GM)
Table 2. Criteria to analyze the most relevant characteristics of conventional and emerging technologies.
Table 2. Criteria to analyze the most relevant characteristics of conventional and emerging technologies.
CharacteristicCriteria
1234
Technological Development LevelIncipientEmergingMedium–highHigh
Operation ModeComplexModerately complexRelatively easyEasy
Characteristic123
Feasibility of Operation with NCRElowmediumhigh
Pretreatment Leveldemandingmoderatesimple
Ease of Industrial ScalingLow*Medium*High*
Table 3. Criteria to analyze the industrial scale-up potential (category 2).
Table 3. Criteria to analyze the industrial scale-up potential (category 2).
ParameterCriteria
00.51
Water quality according to regulationsLow qualityMedium qualityHigh quality
Productivity, quality, and cost vs. ROWorstEqualBest
Innovations in the operationWorst*Equal*Best*
Integration of renewable energyNonePartialYes
Table 4. Summary of desalination technology characteristics (part 1).
Table 4. Summary of desalination technology characteristics (part 1).
TechnologyProcess TypeSeparation GradientPrincipal EquipmentSeparation MechanismEnergy SourceInvestment (USD/m3 d)Electricity Consumption (kWh/m3)
MVCThermal/TraditionalTemperatureThermal compressorsLiquid–vapor equilibriumFossil fuel1000–12005.5–6.5
MEDThermal/TraditionalTemperatureHeat exchanger (multi-effect)Liquid–vapor equilibriumFossil fuel850–20001.5–2.5
MSFThermal/TraditionalTemperatureHeat exchanger (flash)Liquid–vapor equilibriumFossil fuel900–20003.0–4.5
SDThermal/EmergingTemperatureSolar collectorsLiquid–vapor equilibriumSolar radiation500–10000.05
ROMembrane/TraditionalTotal pressureDense membraneSelective permeation. Solution diffusion modelElectricity800–25002.5–3.5
NFMembrane/TraditionalTotal pressureMicroporous membraneIons size vs. pore size
Donnan steric partition pore model (DSPM)
Electricity600–10001.0–2.0
NMMembrane/ModifiedTotal pressureDense and modified membraneSelective permeation improved with nanoparticles Solution diffusion model, share adsorption–desorptionElectricity800–20002.0–3.0
MDMembrane/EmergingPartial pressure and temperatureMesoporous–Macroporous membraneLiquid–vapor equilibriumElectricity<8000.1 (DCMD)
1.0 (VMD)
FOMembrane/EmergingConcentrationDense membraneOsmotic pressureElectricity<8000.15 for each recirculation
GOMembrane/EmergingTotal pressureMonolayer of graphite atoms linked by covalent bondMolecular sieve + adsorption–desorptionElectricity, chemical**
ABMBiomembrane/EmergingConcentrationAquaporin water channels in membraneHydrogen bonds with cell membranesElectricity, chemical**
* The incipient development of these technologies does not allow reporting information at the industrial level.
Table 5. Summary of desalination technology characteristics (part 2).
Table 5. Summary of desalination technology characteristics (part 2).
TechnologyIndication with Regard to Energy ConsumptionOperation Cost (USD/m3)Maintenance Cost (USD/m3)Desalinated Water Quality (ppm)Area/VolumeEffect on Environment
MVCa low enthalpy heat source allows a competitive process0.5–5.00.1010mediumCO2 emissions, local sea temperature increase
MEDa low enthalpy heat source allows a competitive process0.4–5.00.1510mediumCO2 emissions, local sea temperature increase
MSFa low enthalpy heat source allows a competitive process0.4–5.00.1510mediumCO2 emissions, local sea temperature increase
SDoperation with minimal energy consumption0.05–0.200.05–0.1010highlow effects
RO100% dependent on electrical supply0.8–2.00.10300–500lowlocal sea salinity increase
NF100% dependent on electrical supply0.25–0.50.1015,000 (1 stage)lowlocal sea salinity increase
NM100% dependent on electrical supply0.6–1.80.10200–350lowlocal sea salinity increase
MDoperation with low energy consumption1.2–2.50.10<10highlocal sea salinity minimal increase
FOoperation with low energy consumption0.30.10<10mediumlocal sea salinity minimal increase
GOoperation with low energy consumption**<200lowlocal sea salinity increase
ABMoperation with low energy consumption**<200lowlocal sea salinity increase
* The incipient development of these technologies does not allow reporting information at the industrial level.
Table 6. Results of the comparative evaluation of desalination technology characteristics.
Table 6. Results of the comparative evaluation of desalination technology characteristics.
TechnologyTechnological Development LevelOperation ModeFeasibility of Operation with NCREPretreatment LevelEase of Industrial ScalingTotalCompliance Percentage (%)
MVC441211270.6
MED442221482.4
MSF442221482.4
SD333221376.5
RO443131588.2
NF343131482.4
NM343131482.4
MD12221847.1
FO332221270.6
GO11311741.2
ABM11311741.2
Table 7. Results of the comparative evaluation of the industrial scale-up potential for NF, NM, FO + RO, and SD.
Table 7. Results of the comparative evaluation of the industrial scale-up potential for NF, NM, FO + RO, and SD.
ParameterTechnology
NFNMFO + ROSD
Water Quality According to Regulations0111
Productivity, Quality, and Cost vs. RO010.50.5
Innovations in the Operation0.5110.5
Integration of Renewable Energy1111
Total Compliance Percentage (%)1.543.53
37.510087.575
Table 8. Capital expenditure (equipment) for O1, O2, and O3.
Table 8. Capital expenditure (equipment) for O1, O2, and O3.
OptionItemQuantityUnitUSD
O1Solar panel (310 W)80un15,856
Inverter4un4865
Other components 1719
Total (solar park) 22,440
O2Solar panel (310 W)80un25,369
Inverter4un4865
Storage batteries 38,400
Other components 3093
Total (solar park) 65,541
O3Wind turbine (25 kW)25un45,688
Total 45,688
Table 9. Summary of CAPEX of the seawater desalination system.
Table 9. Summary of CAPEX of the seawater desalination system.
ActivityCost (USD)
Seawater desalination system with daily production of 100 m3/d, and a 40 ft container with thermal and acoustic insulation196,442
Potabilization system12,522
Supervision and assembly32,496
Total241,460
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saavedra, A.; Valdés, H.; Mahn, A.; Acosta, O. Comparative Analysis of Conventional and Emerging Technologies for Seawater Desalination: Northern Chile as A Case Study. Membranes 2021, 11, 180. https://doi.org/10.3390/membranes11030180

AMA Style

Saavedra A, Valdés H, Mahn A, Acosta O. Comparative Analysis of Conventional and Emerging Technologies for Seawater Desalination: Northern Chile as A Case Study. Membranes. 2021; 11(3):180. https://doi.org/10.3390/membranes11030180

Chicago/Turabian Style

Saavedra, Aldo, Hugo Valdés, Andrea Mahn, and Orlando Acosta. 2021. "Comparative Analysis of Conventional and Emerging Technologies for Seawater Desalination: Northern Chile as A Case Study" Membranes 11, no. 3: 180. https://doi.org/10.3390/membranes11030180

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