Next Article in Journal
Controllable Liquid Crystal Micro Tube Laser
Previous Article in Journal
As-Grown Domain Structure in Calcium Orthovanadate Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fe2O3 Nanoparticles Deposited over Self-Floating Facial Sponge for Facile Interfacial Seawater Solar Desalination

1
School of Electronic Engineering, Nanjing Xiaozhuang University, Nanjing 211171, China
2
Institute of Quantum Optics and Quantum Information, School of Science, Xi’an Jiaotong University (XJTU), Xi’an 710049, China
3
School of Materials Science and Engineering, Hubei University, Wuhan 430062, China
4
ERC Research Centre, COMSATS University Islamabad, Lahore Campus, Islamabad 54000, Pakistan
5
Industrial Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
6
School of Chemistry & Environment, South China Normal University, Guangzhou 510006, China
7
School of Energy and Environment, Southeast University, Nanjing 210096, China
8
Institute of Science and Technology (IST), 3400 Klosterneuburg, Austria
*
Authors to whom correspondence should be addressed.
Submission received: 3 November 2021 / Revised: 24 November 2021 / Accepted: 29 November 2021 / Published: 3 December 2021
(This article belongs to the Section Hybrid and Composite Crystalline Materials)

Abstract

:
A facile approach for developing an interfacial solar evaporator by heat localization of solar-thermal energy conversion at water-air liquid composed by in-situ polymerization of Fe2O3 nanoparticles (Fe2O3@PPy) deposited over a facial sponge is proposed. The demonstrated system consists of a floating solar receiver having a vertically cross-linked microchannel for wicking up saline water. The in situ polymerized Fe2O3@PPy interfacial layer promotes diffuse reflection and its rough black surface allows Omni-directional solar absorption (94%) and facilitates efficient thermal localization at the water/air interface and offers a defect-rich surface to promote heat localization (41.9 °C) and excellent thermal management due to cellulosic content. The self-floating composite foam reveals continuous vapors generation at a rate of 1.52 kg m−2 h−1 under one 1 kW m−2 and profound evaporating efficiency (95%) without heat losses that dissipates in its surroundings. Indeed, long-term evaporation experiments reveal the negligible disparity in continuous evaporation rate (33.84 kg m−2/8.3 h) receiving two sun solar intensity, and ensures the stability of the device under intense seawater conditions synchronized with excellent salt rejection potential. More importantly, Raman spectroscopy investigation validates the orange dye rejection via Fe2O3@PPy solar evaporator. The combined advantages of high efficiency, self-floating capability, multimedia rejection, low cost, and this configuration are promising for producing large-scale solar steam generating systems appropriate for commercial clean water yield due to their scalable fabrication.

1. Introduction

Freshwater scarcity has emerged as a serious global challenge, which can endanger the survival of all of the species in the hemisphere. Although 70% of our planet’s area is covered by water, unfortunately only 3% comprises freshwater, of which 66% is unavailable due to glaciers and frozen lakes, ice caps, and being underground [1,2]. Ever-increasing industrial growth, urbanization, climate change, and population growth have made the availability of freshwater an unprecedented challenge to mankind [3]. Existing technologies such as membrane-based (reverse osmosis (RO), forward osmosis (FO)), and thermal-based (multistage flashing (MSF), multi-effect distillation (MED)) technologies are widely used for obtaining freshwater; however, the high energy demand and variable feasibility hinder their implementation in poor countries [4,5]. For example, MSF is an energy-intensive technology and is a trade-off in various countries based upon their fuel resources. FO is still premature and cannot be used in the large-scale mass production of freshwater [6,7,8]. To tackle this issue and to envision a more sustainable future, efforts are being made to develop green and renewable technologies for obtaining freshwater that would satisfy the current needs, mainly in those countries. Various innovative technologies are being explored by researchers to meet this global challenge. These include graphene-based membranes [9], nanofiltration (NF) [10], nano adsorbent chillers [11], and solar-driven evaporation [12,13]. These emerging technologies are capable of converting waste heat into useful products in low-temperature conditions in remote sensing areas. Sztekler et al. [14] reported a silica gel-based three-bed sorption chiller for effective desalination in low-temperature conditions.
Solar-driven water evaporation has achieved tremendous interest owing to its excellent heat accumulation on the top surface with perfect thermal insulation towards downward heat flow. In solar stills, the poor optical concentration causes heat generation at the surface of the matrix while vapors are generated elsewhere in the system. This creates inevitable thermal leakage due to the wide gap between the heat and vapor generation points and ultimately results in lower evaporation [15,16,17]. To address these surface heat losses, the volumetric heating approach was implicated using nanofluids which shift the heat-generating point inside the fluid. Although this strategy has been implemented to some extent, it is inefficient for large-scale applications where high surface temperatures are required [18]. An interfacial evaporation strategy was proposed as a solution to sustain heat storage at the air–liquid interface and concurrently reduce heat losses due to radiation, conduction and convection with an output evaporation efficiency of up to ~90%, even under lower optical intensity conditions [19,20]. This approach relies on selective heating of the matrix where water entangles with the photothermal layer and evaporates immediately. This would avoid the volumetric heating which causes the solar-thermal conversion heat to disperse into the entire volume. The possibilities of harvesting sustainable solar energy make solar steam generation technology a promising path in the near future [21]. However, the practical applications of solar steam generating devices still suffer from poor thermal management, mechanical strength, thermal losses, and blockage of steam escape channels, which often occurs due to salt accumulation within the capillary channels for water supply, hence lowering the efficiency of solar-driven water yielding devices [22].
Numerous photothermal materials have been explored to harvest and convert the maximum solar flux into thermal energy. Fe2O3 nanoparticles (NPs) have been reported to have excellent photothermal conversion [23], stability, and propensity for the removal of contaminants from water [24,25]. The valance band edge of Fe2O3 is positive compared to the oxidation potential of water and it can be employed to achieve photocatalytic water oxidation reaction in the vicinity of sacrificial electron acceptors [26,27,28]. The proposed system in this study offers several advantages such as in situ conducting polymerization systems with π-conjugated electron systems [20] that have been recognized for having high absorption coefficients, smooth water transport channels, efficient thermal management, and good environmental stability. However, the fabrication and photothermal investigations of in situ polymerized Fe2O3@PPy-based nanocomposites have not been explored in the literature.
Herein, we demonstrate a solar-empowered interfacial thermal localization inspired Fe2O3@PPy/facial sponge solar evaporation structure (Fe2O3@PPy) which perfectly restricts the heat localization on the heating matrix and forbids downward heat flow. The pitch-black in situ polymerized Fe2O3 photothermal layer enabled outstanding solar absorption (94%), excellent hydrophilicity, and sustainment under intense operating seawater circumstances (3.5 wt.%, no surface degradation/8.3 h continuous operation). An evaporation rate was recorded up to 1.52 kg m−2 h−1 under one sun solar intensity, whereas it is enhanced to 33.84 kg m−2/8.3 h under stimulated seawater conditions. The experiment yields an outstanding solar-thermal conversion efficiency potential of 95%, excluding thermal losses. Besides, the Fe2O3@PPy solar evaporator sustained excellent surface temperature (41.9 °C) under 1 kW m−2. Raman spectroscopy revealed a complete rejection of methyl orange dye solution before and after desalination through a Fe2O3@PPy interfacial solar steam generator. Additionally, Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES) confirmed heavy metal rejection in condensed water which signifies that its NF potential meets the drinking water standard set by the World Health Organization (WHO). This technique lessens the utilization of Fe2O3@PPy as photothermal material while achieving high vapor flux. This offers an opportunity to expand the application of solar-thermal devices in compact, stand-alone, and portable systems.

2. Experimentation

2.1. Materials

All the required chemicals FeCl3.6H2O, NH4OH, and ethanol were acquired from Wuhan BASF Chemical Industries Co., Ltd., Wuhan, China, whereas pyrrole, phytic acid (PA, 50 wt.% in water), and buffered salts were purchased from Wuhan Guangfu fine synthetic industry, Wuhan, China. All the chemicals throughout the study were of standard analytical grade and were processed without further purification.

2.2. Preparation of Fe2O3 Nanoparticles

Fe2O3 NPs were synthesized by the well-known co-precipitation method. For this, iron oxide precursor FeCl3.6H2O (10. g) was dissolved evenly in deionized (DI) water (150 mL) using a magnetic stirrer at room temperature. Then, 2 mL ammonium hydroxide (NH4OH) solution was poured dropwise into the as-prepared solution at the rate of 1 mL min−1 to initiate the precipitation. However, the pH was controlled during the whole synthesis and maintained up to 1. The following black suspension was stirred for 1 h at room temperature until the fine precipitates were obtained. Afterward, the followed suspension was stirred and heated at 80 °C for 2 h to yield a brown powder. The obtained brown powder was allowed to cool freely and was calcined at 500 °C for 4 h to achieve a pure crystalline hexagonal phase.

2.3. Deposition of Fe2O3 NPs over the Facial Sponge

The prepared Fe2O3 NPs were deposited over the circular-shaped facial sponge. Initially, the commercially available facial sponge was crafted into a circular shape with different diameters (3 to 4.5 cm). The Fe2O3 NPs were immersed into a volatile terpineol binder (C10H18O) using a mortar and pestle and a fine slurry was obtained. The slurry was coated evenly over the circular-shaped facial sponge using a BOSOBO paintbrush. After deposition of Fe2O3 NPs, it was preheated using a convection oven at 100 °C for 1 h to evaporate terpineol binder, cooled down freely, and presented for the further in situ polymerization process.

2.4. Synthesis of Polypyrrole Monomers

In situ polymerization of conducting polypyrrole was carried out on Fe2O3 NPs to enhance its omnidirectional solar absorption, NPs stability, and excellent wettability. For in situ polymerization of Fe2O3 NPs coated facial sponge, 2.76 g of ammonium persulfate, (NH4)2 S2O8, was thoroughly dissolved into deionized water (5 mL) and labeled as solution “A”. Afterwards, 5 mL of the isopropanol alcohol (IPA) was taken into another beaker in which 0.86 mL pyrrole was poured along with 1.85 mL phytic acid (1:1 wt.% in water) while stirring continuously until the formation of a homogeneous mixture of dark brownish color which was labeled as solution “B”. Subsequently, the “A” and “B” solutions were poured into spray bottles and sprayed on the Fe2O3 -coated facial sponge according to an “A-B-A-B” sequence. The successive repetition of the spraying sequence was followed until the formation of a pitch dark, homogeneous coating of polypyrrole was accomplished on the Fe2O3/facial sponge surface, as illustrated in Figure 1. The deposited mass of polypyrrole on the Fe2O3/facial sponge was calculated as 28 mg by estimating the comparative change in mass strategy. Detailed characterization is given in Supporting Information S1.

2.5. Solar to Vapor Conversion Efficiency (η)

The solar to vapor conversion efficiency (η) of the Fe2O3@PPy evaporation system was calculated using Equations (1) and (2) [8,29]:
η = h L V q i
h L V = λ + C Δ T
For the above equation, denotes the evaporation rate of the system when exposed to solar irradiation intensity subtracting the evaporation rate for bulk water (mass flux) and in the absence of light, h L V denotes the phase change enthalpy from liquid-vapor along with sensible heat, and q i is the irradiated solar intensity (1 kW m−2). λ denotes the phase-changing latent heat (2430 kJ kg−1 at 30 °C and 2256 kJ kg−1 at 100 °C), C is the constant value which is 4.2 kJ kg−1 K−1 (water-specific heat capacity), and Δ T gives the elevation in the temperature of the water. The approximate ambient temperature (24.3 °C) and humidity (43%) conditions were recorded.

2.6. Energy Balance (Heat Losses)

The energy balance equations for the Fe2O3@PPy solar evaporator were explained by thermal transport theory under one sun intensity (1 kW m−2) [18,30]. The photothermal conversion efficiency of the Fe2O3@PPy solar evaporator was evaluated (Equations (3)–(5)) by assessing conductive heat losses via the following three phenomena: heat losses to bulk water via conduction (QConduction), convective heat losses to the surroundings (QConvection), and radiative heat losses into the air (QRadiation):
For conductive heat losses:
Q C o n d u c t i o n = A k T 1   T 2 Δ l
where A denotes the cross-section surface area of the device and k shows thermal conductivity to the bulk water (0.6 W m−1 K−1) [31]. The water temperature was recorded at two points via two embedded thermocouples separated by intermediate distance (∆l).
For convective heat losses:
Q c o n v e c t i o n = h   T s   T v
where h is the convective heat transport coefficient (approximately 10 W m−2 K−1), T s is for the top surface temperature of solar-thermal conversion device, and T v denotes the rise in the surrounding environmental temperature near the top surface due to the generation of hot vapors.
For radiative heat losses:
Q R a d i a t i o n = ε   σ   T s   4 T   4
where ε is the emission of the absorber (~0.93) [32], σ is the Stefan–Boltzmann constant (5.669 × 10−8 W m−2 K−4), and T records the temperature of the adjacent surroundings. Upon the generation of hot vapors, the top surface which is covered by the vapors appears to be semitransparent to solar irradiation; as a result, heat loss can be determined by maximum and minimum values of radiative heat loss recorded as T = T V (vapor temperature) and T = T a (surrounding temperature), respectively.

2.7. Solar-Driven Steam Generation Setup

Solar-driven evaporation experiments were executed under a solar simulator (PLS-FX300HU) outputting the constant incident intensity at 1 kW m−2 (1 sun). A standard spectrum of 1.5 G AM was acquired by employing an optical filter. The self-floating and hydrophilic Fe2O3@PPy solar evaporation structure having 25 mm thickness was placed on the water surface filled in a petty dish (simulated seawater, heavy metal contaminated, and methyl orange solution) and was exposed under 1 kW m−2 solar intensity. An advanced electronic balance (Mettler Toledo, ME204) up to 0.001 g resolution was utilized to calculate the time-dependent mass change variation. After the stabilization of the whole evaporation system, all the evaporation rates were measured under one solar intensity. Moreover, the ionic concentrations of salt and heavy metal ions were recorded by using an inductively coupled plasma-optical emission spectrometry (ICP-AES, E.P. Optimal 8000) before and after treating the water. All the experimental measurements were done at ambient environmental conditions, at closest mean values ~24.3 °C temperature and ~43% humidity. Surface temperatures were recorded by employing the Hand-Held Optical Meter Model.

3. Results and Discussion

3.1. Chemical States and Crystal Structure

Hematite (Fe2O3) is the most attractive and significant metal oxide due to well-visible light absorption, solar energy conversion, environmental compatibility, and excellent chemical stability under intense conditions. It has been used widely for wastewater treatment and photocatalytic applications [33]. Figure 2a presents the structural configuration of the hematite crystalline hexagonal structure. It is an n-type semiconductor that occupies the largest fraction of visible light, the valance band of hematite consists of O 2p and Fe 3d orbitals, constituting a complex structure, and the conductions band consists of 3d orbitals of Fe [33]. Figure 2b shows the X-ray diffraction spectra of Fe2O3 nanoparticles. The XRD patterns of Fe2O3 nanoparticles match perfectly with the standard data (JCPDS 33-0664). The crystal structure analysis shows the hexagonal structure with peaks appearing at 2θ values of 25°, 34°, 36°, 42°, 48°, 54.5°, 57.6°, 62.5°, and 64.1° correspond to the (0 1 2), (1 1 0), (1 1 3), (2 0 2), (0 2 4), (1 1 6), (0 1 8), (2 1 4), and (3 0 0) planes of the Fe2O3 nanoparticles, respectively. No extra peaks appeared in a given pattern, showing its hydroxides transformation to hematite during the coprecipitation process [25].
The XPS survey spectrum was analyzed to study the elemental electronic states present on the surface of Fe2O3 nanoparticles. The homogeneous elemental distribution impacts the oxidation states shifting which could optimize the material’s optical, thermal, and magnetic properties. The XPS survey spectrum of Fe2O3 shows the presence of Fe, O, and C on the sample surface as shown in Figure 2c. The prominent peak appears at 534.44 eV which attributes to the 1s O−2 ions through which the highest range at 710.55 eV corresponds to Fe 2p ions. The deconvolution of the peaks for each element was done by applying Gaussian fit to observe the changes in the homogeneous ion distribution in the Fe2O3 lattice. Figure S1 shows the Fe 2p spectra with ionic states Fe 2p1/2 and Fe 2p3/2 corresponding to 714.15 and 728.55 eV binding energies, respectively. Figure 2d shows the C 1s spectra which have three prominent peaks at binding energies 283, 284.5, and 288 eV, which correlate to C-C, C-O, and O-C=O bonding, respectively. The characteristic bands appearing at 532 and 529 eV are attributed to the O 1s XPS spectra which correspond to the C-O and Fe-O bonding (Figure 2e). The physicochemical characteristics of the prepared Fe2O3 NPs were investigated through the FTIR spectrum, as shown in Figure 2f. The absorption peaks appearing at 1000 and 1300 cm−1 correspond to the Fe-O and Fe-OH stretching vibrations, respectively. Besides, the absorption band appearing at 1600 and 3500 cm−1 are referred to as the presence of hydroxyl group –OH.

3.2. Fabrication of Fe2O3@PPy Solar Evaporator

Figure 3a presents the digital image of the step-by-step preparation of the Fe2O3@PPy-based solar evaporation device. Initially, the super hydrophilic and mechanically robust pristine facial sponge was chosen as substrate and carved carefully in a circular shape with the dimensions of 2 cm × 2 cm × 1 cm = 4 cm3, incorporated by vertically oriented water channels for water transport. This facial sponge has several advantages when compared to commercial polyurethane (PU) foam including flexibility, self-floating ability, and remarkable heat-insulating capability [34]. The obtained reddish-orange color gel was deposited uniformly on a super hydrophilic circular-shaped facial sponge using a BOSOBO paintbrush. Polypyrrole is an organic polymer produced by the oxidative polymerization of pyrrole. It is an intrinsically conducting polymer and has a heterocyclic structure with the formula H(C4H2NH)nH. The surface of polypyrrole films presents fractal properties allowing ionic diffusion with anomalous diffusion patterns [35]. Indeed, polypyrrole coating offers good absorbing capacity, mechanical robustness, biocompatibility, and long-term cytotoxicity [36].
To increase the dispersion and distribution of incident solar radiation (lowering the refractive index) and solar absorbing capacity, a highly conductive polypyrrole polymer was coated on the Fe2O3 surface by following a facile in situ polymerization. For this, the composing of the two monomer solutions is accomplished (as explained above in Section 2.3) and sequentially sprayed on the surface of the Fe2O3/facial sponge until the reddish-orange surface is turned into a pitch dark color. The digital photograph of the Fe2O3@PPy prototype in a different size is presented in Figure 3a. Figure 3b–d demonstrates the FESEM images of Fe2O3 NPs, showing the uniform morphology of the prepared Fe2O3 NPs with an average size of 40 nm, thus having a high surface-to-volume ratio for enhanced photothermal conversion activity. Fe2O3 NPs EDS analysis confirms the presence of Fe (45%), O (45%), and C (10%) as shown in Figure 3f–i, respectively.

3.3. Superhydrophilic Evaporation Channels

Super hydrophilic solar-driven interfacial evaporation systems with an inter-linked porous structure enhance solar to vapor conversion efficiency due to surface roughness with optimized defect chemistry, which ultimately increases the absorption capacity and diffuse reflection [37,38]. Figure 4a,b demonstrates the FESEM images of cross-linked open porous morphology of super hydrophilic facial sponge with a uniform porous structure having plenty of micrometer-sized open-pore channels. The vertical cross-linked water channels occupied by open pores reveal the great potential for capillary pumping of water and a high evaporation rate at the water surface. Figure 4c,d reveals the morphological analysis of the surface configuration of the polymerized Fe2O3@PPy interfacial heating device, showing an inset of complex surface structure where the built-in cross-linked facial sponge surface is coated with a pitch dark color layer of Fe2O3@PPy NPs. The Fe2O3@PPy NPs are uniformly distributed over the surface with no structural alteration, exhibiting an overall condensed texture due to successive polymerizations. The complex surface structure induces multiple rays within the matrix, enabling omnidirectional absorption while offering a large surface area for heat distribution. This would ultimately enhance the light absorption ability of the device while restricting the downward thermal conduction to achieve a high evaporation surface temperature [39]. The in situ polymerization plays a decisive role in the defect chemistry with enhanced solar absorption, wettability, and tortuosity of the device [35,39]. The observed surface configuration and device morphology owe excellent capability for outstanding heat accumulations, smooth water transportation, and prevention of downward thermal conduction to avoid heat losses.

3.4. Fe2O3@PPy Solar Evaporator

The interfacial evaporation devices for seawater desalination reveal the great potential for the maximum utilization of solar energy and controlling energy losses [40]. Figure 5a shows the energy balance for the interfacial Fe2O3@PPy solar evaporator based on interfacial heating which exhibits negligible heat losses under solar irradiation. It also accompanies the simultaneous liquid to vapor phase transition by sustainment of heat at the air–liquid interface. The ultra-black composite surface effectively harvests the incident light energy for multiple diffuse reflections which optimize the incident light energy and control light losses.
The incoming solar energy is typically divided into two components, solar to vapor efficiency and undesired heat losses. The overall efficiency of the interfacial Fe2O3@PPy solar evaporator was evaluated (95%) excluding the heat losses to the environment, e.g., thermal loss via conduction to underlying water (3.1%), heat transfer via convection through vapors (1.2%), and thermal loss to the adjacent environment via radiation (0.7%). The optical absorptions of the Fe2O3 and Fe2O3@PPy solar evaporator were analyzed by UV-Vis spectroscopy over the whole solar spectrum (200–2500 nm). The UV-Vis absorption spectrum of pure Fe2O3 and in situ polymerized Fe2O3@PPy solar evaporator are demonstrated in Figure 5b. An outstanding solar absorption (94%) potential is exhibited by the in situ polymerized Fe2O3@PPy interfacial solar evaporator and is significantly enhanced compared to the simple Fe2O3 NPs with negligible transmission (4%) and reflection (2%) calculated by the following absorption formula: (1-T-R).
The supreme solar-harvesting potential can be featured in the crimped structural configuration of in situ polymerizations of Fe2O3 NPs which can efficiently sustain and localize the heat by intrinsically dispensing the incoming solar radiations on the composite surface, which sculpts the basic mechanism of the interfacial solar evaporating structure [41]. Herein, we demonstrate a comparative analysis of four systems for the interfacial solar evaporation phenomena, including pure water, super hydrophilic facial sponge, Fe2O3/facial sponge, and Fe2O3@PPy solar evaporator. Each of the evaporation systems was exposed under simulated one solar irradiation (1 kW m−2) and time-dependent mass change was recorded via a weighing balance with 0.0001 g resolution for one hour. An enhancement in thermal localization over the top surface escalates the vapor generation; thus, the Fe2O3@PPy solar steam generator shows the outstanding mass variations (1.52 kg m−2) comparative to the bulk water and other evaporation systems, as shown in Figure 5c. Furthermore, mass change for Fe2O3@PPy solar evaporator for multiple solar intensities was recorded, as provided in Figure S2. Although there are some solar steam generation devices reported for good evaporation rate, the major challenge being faced by these devices is the deformation of structural configuration with mechanical frailty mainly due to salt accumulation and surface distortion by the deposition of organic and inorganic impurities present in seawater, which leads to the compositional disintegration after operating a few cycles and lower efficiency of the device [5,12]. To address this, the evaporating device was operated for continuous operation for long-term performance to inspect the mechanical strength and persistence in evaporation efficiency, which reveals that Fe2O3@PPy remarkably achieved the consistency in steam generation under simulated seawater for continuous 8.3 h (33.84 kg m−2) with the minimal disparity in the evaporation rate under 1 kWm−2, showing excellent scalability for operating over a long time period (Figure 5d). The comparative analysis of temperature gradient for the bulk water, vapor temperature, and Fe2O3@PPy solar evaporating device under one solar intensity (1 kW m−2) is illustrated in Figure 5e. The surface temperature achieved a maximum value up to 41.9 °C under one sun illumination by a Fe2O3@PPy solar evaporator that is comparatively more efficient than other solar-thermal conversion devices. This is attributed to the remarkable heat localization on the photothermal layer and excellent insulation from downward heat conduction which ultimately leads to achieving higher evaporation efficiency. Figure 5f demonstrates the analysis of the comparative evaporation rate for the four developed systems, which reveals that the evaporation rate of the Fe2O3@PPy solar evaporator (1.52 kg m−2 h−1) and photothermal conversion efficiency (95%) are notably higher comparative to the simple Fe2O3/facial sponge and many other evaporation systems to date [42,43,44]. However, a detailed solar evaporation efficiencies-based comparison is provided in Supplementary Materials Table S1.

3.5. Self-Rejection Potential

The self-floatable Fe2O3@PPy solar evaporator absorbs the entire solar spectrum and is transformed into confined thermal packs for efficient vapor generation which can be condensed to pure freshwater by completely rejecting heavy metal ions, primary salt ions, and wastewater contamination. Stimulated seawater and heavy metal ions, as well as methyl orange dye solutions were prepared to verify the self-rejection potential of the Fe2O3@PPy solar evaporator. More importantly, Raman spectroscopy analysis of methyl orange dye solution was carried out before evaporation and in condensed water. Figure 6a,b represent the microscopical image of the methyl orange dissolved water solution, and the condensed water exhibiting the complete removal of methyl orange and pure water droplets could be seen clearly. Figure 6c represents the Raman spectra of methyl orange dissolved water solution and condensed water after treating through interfacial Fe2O3@PPy solar evaporator. The methyl orange dissolved solutions show the methyl orange band appearing at 1500 and 1600 cm−1, exhibiting the presence of methyl orange, and condensed water shows a smooth spectrum with no bands appearing in the range of 1400 to 1600 cm−1, which reveals the complete removal of methyl orange after evaporation through the interfacial Fe2O3@PPy solar evaporator.
The Fe2O3@PPy solar evaporator encompasses the excellent self-resistant capability against the salt and heavy metal ions within the vertically oriented cross-linked macro water channels of super hydrophilic facial sponge and ensures the standard of purity level of freshwater yield. As the cross-linked water transportation network promotes water evaporation only through the molecular mode, it also assists in water purification against the removal of primary metal ions and heavy metal ions. ICP-AES was employed to assess the soluble salt ions concentration gradient (3.5 wt.% NaCl, CaCl2, KCl, MgSO4), and the desalination potential of the Fe2O3@PPy solar evaporator was measured through simulated and condensed water. The result obtained from ICP-AES exhibits an appreciable decrease in the ionic concentrations of Na+, K+, Ca2+, and Mg2+ in the collected water comparative to the initial concentration in the simulated seawater. The result is far below the standards of drinking water set by the WHO [29,45], thus fetching the application of Fe2O3@PPy solar evaporator up to standard drinking water as illustrated in Figure 6d. Furthermore, the commercial waste contains high concentrations of heavy metal ions which are hazardous and predominately contaminate the water [46,47]. The condensed water was also examined for the engrossment of intruded heavy metals ions, e.g., Cu2+, Fe3+, and Co2+; the concentration level of these heavy metal ions was significantly lessened from 1000 mg L−1 to 0.0084, 0.0065, and 0.0032 mg L−1 after evaporation via the Fe2O3@PPy solar evaporator as shown in Figure 6e. The self-resistant potential of the interfacial Fe2O3@PPy solar evaporator against rejection of primary metals ion and heavy metal ions of simulated seawater to purify the water meets 100% of the standards for drinking water level as demonstrated in Figure 6f, proving the outstanding potential for treating sewage water for freshwater production.

4. Conclusions

In summary, a self-floating and flexible in situ polymerized Fe2O3@PPy solar evaporator was composed of a super hydrophilic facial sponge with vertically cross-linked water channels as a solar-driven interfacial evaporating system for intensified heat accumulation and enhanced solar-thermal conversion efficiency. The pitch-dark rough surface texture of the photothermal layer induces multiple rays within the heating matrix, which facilitates the omnidirectional solar absorption and offers defect abundant surface to promote heat localization and achieve high surface temperature (41.9 °C) along with high evaporation efficiency (90% excluding heat losses). The self-dissolving potential against methyl-orange and multimedia rejection exhibit an excellent functioning capability and self-regenerating potential against various impurities while maintaining the structural configuration without any deformation or antifouling nature. The presented work paves the way for new prospects for engineering a low cost, highly efficient with excellent hat localization, scalable, easily manufactured, and industrially compatible interfacial solar evaporator with commercially available precursors having excellent potential against the rejection of various impurities. Moreover, future work on the determinization of mechanical resistance of the system with regards to relatively prolonged operating cycles can help to quantify the operational cost and maintenance cost of the system. It also makes our device a promising solution for implementing at an industrial scale in remote locations to cope with water scarcity issues.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/cryst11121509/s1, Note S1: Characterization; Figure S1. Fe 2p scan of Fe2O3 NPs; Figure S2. Time dependent mass change under different solar intensities; Table S1. Detailed performance comparison with the recent successful solar evaporators.

Author Contributions

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

Funding

The APC was funded by Researchers Supporting Project number (RSP-2021/387), King Saud University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to King Saud University for funding this work through Researchers Supporting Project number (RSP-2021/387), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Wu, X.; Wu, Z.; Wang, Y.; Gao, T.; Li, Q.; Xu, H. All-Cold Evaporation under One Sun with Zero Energy Loss by Using a Heatsink Inspired Solar Evaporator. Adv. Sci. 2021, 8, 2002501. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, S.; Fan, Y.; Wang, F.; Su, Y.; Zhou, X.; Zhu, Z.; Sun, H.; Liang, W.; Li, A. Potentially scalable fabrication of salt-rejection evaporator based on electrogenerated polypyrrole-coated nickel foam for efficient solar steam generation. Desalination 2021, 505, 114982. [Google Scholar] [CrossRef]
  3. Tao, P.; Ni, G.; Song, C.; Shang, W.; Wu, J.; Zhu, J.; Chen, G.; Deng, T. Solar-driven interfacial evaporation. Nat. Energy 2018, 3, 1031–1041. [Google Scholar] [CrossRef]
  4. Awasthi, A.; Kumari, K.; Panchal, H.; Sathyamurthy, R. Passive solar still: Recent advancements in design and related performance. Environ. Technol. Rev. 2018, 7, 235–261. [Google Scholar] [CrossRef]
  5. Wang, P. Emerging investigator series: The rise of nano-enabled photothermal materials for water evaporation and clean water production by sunlight. Environ. Sci. Nano 2018, 5, 1078–1089. [Google Scholar] [CrossRef] [Green Version]
  6. Liu, X.; Mishra, D.D.; Wang, X.; Peng, H.; Hu, C. Towards highly efficient solar-driven interfacial evaporation for desalination. J. Mater. Chem. A 2020, 8, 17907–17937. [Google Scholar] [CrossRef]
  7. Shi, Y.; Zhang, C.; Li, R.; Zhuo, S.; Jin, Y.; Shi, L.; Hong, S.; Chang, J.; Ong, C.; Wang, P. Solar evaporator with controlled salt precipitation for zero liquid discharge desalination. Environ. Sci. Technol. 2018, 52, 11822–11830. [Google Scholar] [CrossRef]
  8. Xu, J.; Wang, Z.; Chang, C.; Fu, B.; Tao, P.; Song, C.; Shang, W.; Deng, T. Solar-driven interfacial desalination for simultaneous freshwater and salt generation. Desalination 2020, 484, 114423. [Google Scholar] [CrossRef]
  9. Homaeigohar, S.; Elbahri, M. Graphene membranes for water desalination. NPG Asia Mater. 2017, 9, e427. [Google Scholar] [CrossRef] [Green Version]
  10. Dayarathne, H.N.P.; Choi, J.; Jang, A. Enhancement of cleaning-in-place (CIP) of a reverse osmosis desalination process with air micro-nano bubbles. Desalination 2017, 422, 1–4. [Google Scholar] [CrossRef]
  11. Sztekler, K.; Kalawa, W.; Nowak, W.; Mika, Ł.; Krzywański, J.; Grabowska, K.; Sosnowski, M.; Alharbi, A.A. Performance Evaluation of a Single-Stage Two-Bed Adsorption Chiller with Desalination Function. J. Energy Resour. Technol. 2020, 143, 082101. [Google Scholar] [CrossRef]
  12. Irshad, M.S.; Arshad, N.; Wang, X. Nanoenabled Photothermal Materials for Clean Water Production. Glob. Chall. 2021, 5, 2000055. [Google Scholar] [CrossRef]
  13. Fuzil, N.S.; Othman, N.H.; Alias, N.H.; Marpani, F.; Othman, M.H.D.; Ismail, A.F.; Lau, W.J.; Li, K.; Kusworo, T.D.; Ichinose, I.; et al. A review on photothermal material and its usage in the development of photothermal membrane for sustainable clean water production. Desalination 2021, 517, 115259. [Google Scholar] [CrossRef]
  14. Sztekler, K.; Kalawa, W.; Nowak, W.; Mika, L.; Gradziel, S.; Krzywanski, J.; Radomska, E. Experimental Study of Three-Bed Adsorption Chiller with Desalination Function. Energies 2020, 13, 5827. [Google Scholar] [CrossRef]
  15. Ito, Y.; Tanabe, Y.; Han, J.; Fujita, T.; Tanigaki, K.; Chen, M. Multifunctional porous graphene for high-efficiency steam generation by heat localization. Adv. Mater. 2015, 27, 4302–4307. [Google Scholar] [CrossRef]
  16. He, S.; Chen, C.; Kuang, Y.; Mi, R.; Liu, Y.; Pei, Y.; Kong, W.; Gan, W.; Xie, H.; Hitz, E.; et al. Nature-inspired salt resistant bimodal porous solar evaporator for efficient and stable water desalination. Energy Environ. Sci. 2019, 12, 1558–1567. [Google Scholar] [CrossRef]
  17. Irshad, M.S.; Wang, X.; Abbas, A.; Yu, F.; Li, J.; Wang, J.; Mei, T.; Qian, J.; Wu, S.; Javed, M.Q. Salt-resistant carbon dots modified solar steam system enhanced by chemical advection. Carbon 2021, 176, 313–326. [Google Scholar] [CrossRef]
  18. Ghasemi, H.; Ni, G.; Marconnet, A.M.; Loomis, J.; Yerci, S.; Miljkovic, N.; Chen, G. Solar steam generation by heat localization. Nat. Commun. 2014, 5, 4449. [Google Scholar] [CrossRef] [Green Version]
  19. Yin, H.; Xie, H.; Liu, J.; Zou, X.; Liu, J. Graphene tube shaped photothermal layer for efficient solar-driven interfacial evaporation. Desalination 2021, 511, 115116. [Google Scholar] [CrossRef]
  20. Irshad, M.S.; Wang, X.; Abbasi, M.S.; Arshad, N.; Chen, Z.; Guo, Z.; Yu, L.; Qian, J.; You, J.; Mei, T. Semiconductive, Flexible MnO2 NWs/Chitosan Hydrogels for Efficient Solar Steam Generation. ACS Sustain. Chem. Eng. 2021, 9, 3887–3900. [Google Scholar] [CrossRef]
  21. Liu, J.; Zou, X.; Cai, Z.; Peng, Z.; Xu, Y. Polymer based phase change material for photo-thermal utilization. Sol. Energy Mater. Sol. Cells 2021, 220, 110852. [Google Scholar] [CrossRef]
  22. Li, X.; Cooper, T.; Xie, W.; Hsu, P.-C. Design and Utilization of Infrared Light for Interfacial Solar Water Purification. ACS Energy Lett. 2021, 6, 2645–2657. [Google Scholar] [CrossRef]
  23. Larsen, G.K.; Farr, W.; Hunyadi Murph, S.E. Multifunctional Fe2O3–Au Nanoparticles with Different Shapes: Enhanced Catalysis, Photothermal Effects, and Magnetic Recyclability. J. Phys. Chem. C 2016, 120, 15162–15172. [Google Scholar] [CrossRef]
  24. Mansourpanah, Y.; Rahimpour, A.; Tabatabaei, M.; Bennett, L. Self-antifouling properties of magnetic Fe2O3/SiO2-modified poly (piperazine amide) active layer for desalting of water: Characterization and performance. Desalination 2017, 419, 79–87. [Google Scholar] [CrossRef]
  25. Gopal, R.A.; Song, M.; Yang, D.; Lkhagvaa, T.; Chandrasekaran, S.; Choi, D. Synthesis of hierarchically structured ɤ-Fe2O3–PPy nanocomposite as effective adsorbent for cationic dye removal from wastewater. Environ. Pollut. 2020, 267, 115498. [Google Scholar] [CrossRef] [PubMed]
  26. Cao, Y.-Q.; Zi, T.-Q.; Zhao, X.-R.; Liu, C.; Ren, Q.; Fang, J.-B.; Li, W.-M.; Li, A.-D. Enhanced visible light photocatalytic activity of Fe2O3 modified TiO2 prepared by atomic layer deposition. Sci. Rep. 2020, 10, 13437. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, Y.; Sun, N.; Hu, J.; Li, S.; Qin, G. Photocatalytic degradation properties of α-Fe2O3 nanoparticles for dibutyl phthalate in aqueous solution system. R. Soc. Open Sci. 2018, 5, 172196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Liu, B.; Wang, D.; Li, H.; Xu, Y.; Zhang, L. As(III) removal from aqueous solution using α-Fe2O3 impregnated chitosan beads with As(III) as imprinted ions. Desalination 2011, 272, 286–292. [Google Scholar] [CrossRef]
  29. Arshad, N.; Ahmed, I.; Irshad, M.S.; Li, H.R.; Wang, X.; Ahmad, S.; Sharaf, M.; Firdausi, M.; Zaindin, M.; Atif, M. Super Hydrophilic Activated Carbon Decorated Nanopolymer Foam for Scalable, Energy Efficient Photothermal Steam Generation, as an Effective Desalination System. Nanomaterials 2020, 10, 2510. [Google Scholar] [CrossRef]
  30. Kabeel, A.E.; El-Agouz, S.A. Review of researches and developments on solar stills. Desalination 2011, 276, 1–12. [Google Scholar] [CrossRef]
  31. Schön, S.J. Chapter 9—Thermal properties. In Physical Properties of Rocks; Schön, J.H., Ed.; Handbook of Petroleum Exploration and Production; Elsevier: Amsterdam, The Netherlands, 2011; Volume 8, pp. 337–372. [Google Scholar]
  32. Wang, P.; Gu, Y.; Miao, L.; Zhou, J.; Su, H.; Wei, A.; Mu, X.; Tian, Y.; Shi, J.; Cai, H. Co3O4 nanoforest/Ni foam as the interface heating sheet for the efficient solar-driven water evaporation under one sun. Sustain. Mater. Technol. 2019, 20, e00106. [Google Scholar] [CrossRef]
  33. Xue, Y.; Wang, Y. A review of the α-Fe2O3 (hematite) nanotube structure: Recent advances in synthesis, characterization, and applications. Nanoscale 2020, 12, 10912–10932. [Google Scholar] [CrossRef]
  34. Chen, Q.; Pei, Z.; Xu, Y.; Li, Z.; Yang, Y.; Wei, Y.; Ji, Y. A durable monolithic polymer foam for efficient solar steam generation. Chem. Sci. 2018, 9, 623–628. [Google Scholar] [CrossRef] [Green Version]
  35. Xiao, C.; Chen, L.; Mu, P.; Jia, J.; Sun, H.; Zhu, Z.; Liang, W.; Li, A. Sugarcane-Based Photothermal Materials for Efficient Solar Steam Generation. ChemistrySelect 2019, 4, 7891–7895. [Google Scholar] [CrossRef]
  36. Huang, W.; Hu, G.; Tian, C.; Wang, X.; Tu, J.; Cao, Y.; Zhang, K. Nature-inspired salt resistant polypyrrole–wood for highly efficient solar steam generation. Sustain. Energy Fuels 2019, 3, 3000–3008. [Google Scholar] [CrossRef]
  37. Yu, F.; Chen, Z.; Guo, Z.; Irshad, M.S.; Yu, L.; Qian, J.; Mei, T.; Wang, X. Molybdenum Carbide/Carbon-Based Chitosan Hydrogel as an Effective Solar Water Evaporation Accelerator. ACS Sustain. Chem. Eng. 2020, 8, 7139–7149. [Google Scholar] [CrossRef]
  38. Yu, F.; Guo, Z.; Xu, Y.; Chen, Z.; Irshad, M.S.; Qian, J.; Mei, T.; Wang, X. Biomass-Derived Bilayer Solar Evaporator with Enhanced Energy Utilization for High-Efficiency Water Generation. ACS Appl. Mater. Interfaces 2020, 12, 57155–57164. [Google Scholar] [CrossRef]
  39. Xu, Y.; Wang, J.; Yu, F.; Guo, Z.; Cheng, H.; Yin, J.; Yan, L.; Wang, X. Flexible and Efficient Solar Thermal Generators Based on Polypyrrole Coated Natural Latex Foam for Multimedia Purification. ACS Sustain. Chem. Eng. 2020, 8, 12053–12062. [Google Scholar] [CrossRef]
  40. Parsa, S.M.; Rahbar, A.; Koleini, M.H.; Aberoumand, S.; Afrand, M.; Amidpour, M. A renewable energy-driven thermoelectric-utilized solar still with external condenser loaded by silver/nanofluid for simultaneously water disinfection and desalination. Desalination 2020, 480, 114354. [Google Scholar] [CrossRef]
  41. Zong, L.; Li, M.; Li, C. Intensifying solar-thermal harvest of low-dimension biologic nanostructures for electric power and solar desalination. Nano Energy 2018, 50, 308–315. [Google Scholar] [CrossRef]
  42. Wang, C.; Wang, J.; Li, Z.; Xu, K.; Lei, T.; Wang, W. Superhydrophilic porous carbon foam as a self-desalting monolithic solar steam generation device with high energy efficiency. J. Mater. Chem. A 2020, 8, 9528–9535. [Google Scholar] [CrossRef]
  43. Qin, D.-D.; Zhu, Y.-J.; Yang, R.-L.; Xiong, Z.-C. A salt-resistant Janus evaporator assembled from ultralong hydroxyapatite nanowires and nickel oxide for efficient and recyclable solar desalination. Nanoscale 2020, 12, 6717–6728. [Google Scholar] [CrossRef] [PubMed]
  44. Shan, X.; Lin, Y.; Zhao, A.; Di, Y.; Hu, Y.; Guo, Y.; Gan, Z. Porous reduced graphene oxide/nickel foam for highly efficient solar steam generation. Nanotechnology 2019, 30, 425403. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, L.; Li, N.; Guo, C.; He, J.; Wang, S.; Qiao, L.; Li, F.; Yu, L.; Wang, M.; Xu, X. Marine biomass-derived composite aerogels for efficient and durable solar-driven interfacial evaporation and desalination. Chem. Eng. J. 2020, 417, 128051. [Google Scholar] [CrossRef]
  46. Parsa, S.M.; Rahbar, A.; Koleini, M.H.; Davoud Javadi, Y.; Afrand, M.; Rostami, S.; Amidpour, M. First approach on nanofluid-based solar still in high altitude for water desalination and solar water disinfection (SODIS). Desalination 2020, 491, 114592. [Google Scholar] [CrossRef]
  47. Irshad, M.S.; Arshad, N.; Wang, X.; Li, H.R.; Javed, M.Q.; Xu, Y.; Alshahrani, L.A.; Mei, T.; Li, J. Intensifying solar interfacial heat accumulation for clean water generation excluding heavy metal ions, and oil emulsions. Sol. RRL 2021, 5, 2100427. [Google Scholar] [CrossRef]
Figure 1. Schematic demonstration of the synthesis of Fe2O3 NPs, deposition of Fe2O3 NPs over the two-dimensional facial sponge, and it’s in situ polymerization of Fe2O3@PPy interfacial solar steam generator.
Figure 1. Schematic demonstration of the synthesis of Fe2O3 NPs, deposition of Fe2O3 NPs over the two-dimensional facial sponge, and it’s in situ polymerization of Fe2O3@PPy interfacial solar steam generator.
Crystals 11 01509 g001
Figure 2. (af) Crystal structure and chemical states as-synthesized Fe2O3 nanoparticles (NPs); (a) crystallographic unit cell of Fe2O3; (b) XRD spectra of Fe2O3 NPs; (c) XPS survey of Fe2O3 NPs; (d) C 1s spectrum; (e) O 1s spectra of Fe2O3 NPs; (f) FTIR stretching bonds peaks of Fe2O3 NPs.
Figure 2. (af) Crystal structure and chemical states as-synthesized Fe2O3 nanoparticles (NPs); (a) crystallographic unit cell of Fe2O3; (b) XRD spectra of Fe2O3 NPs; (c) XPS survey of Fe2O3 NPs; (d) C 1s spectrum; (e) O 1s spectra of Fe2O3 NPs; (f) FTIR stretching bonds peaks of Fe2O3 NPs.
Crystals 11 01509 g002
Figure 3. (a) Photograph of the step-by-step preparation of Fe2O3@PPy solar evaporator with various sizes. (bf) Fe2O3 nanoparticles under different resolutions. (fi) EDS mapping of Fe2O3 nanoparticles.
Figure 3. (a) Photograph of the step-by-step preparation of Fe2O3@PPy solar evaporator with various sizes. (bf) Fe2O3 nanoparticles under different resolutions. (fi) EDS mapping of Fe2O3 nanoparticles.
Crystals 11 01509 g003
Figure 4. Super hydrophilic and crosslinked water evaporation channels. (a,b) FESEM image of the crosslinked open porous structure of super hydrophillic facial sponge. (c,d) FESEM image of in-situ polymerized Fe2O3/facial sponge showing rugged surface texture with intensified pitch-black color with interlinked vertical channels.
Figure 4. Super hydrophilic and crosslinked water evaporation channels. (a,b) FESEM image of the crosslinked open porous structure of super hydrophillic facial sponge. (c,d) FESEM image of in-situ polymerized Fe2O3/facial sponge showing rugged surface texture with intensified pitch-black color with interlinked vertical channels.
Crystals 11 01509 g004
Figure 5. (a) Energy balance diagram of Fe2O3@PPy solar evaporator with conductive, convective, and radiative losses. (b) Full range of the UV-Vis absorption spectrum of Fe2O3 and Fe2O3@PPy solar evaporator. (c) Mass change in water, facial sponge, Fe2O3/facial sponge, and Fe2O3@PPy solar evaporators under one sun illumination. (d) The long-term evaporation performance with minor disparity under one sun irradiation for continuous evaporation up to 8.3 h in seawater conditions. (e) Surface temperatures of Fe2O3@PPy solar evaporator and vapors. (f) The achieved evaporation rates and respective solar to vapor conversion efficiencies of the designed four evaporation systems.
Figure 5. (a) Energy balance diagram of Fe2O3@PPy solar evaporator with conductive, convective, and radiative losses. (b) Full range of the UV-Vis absorption spectrum of Fe2O3 and Fe2O3@PPy solar evaporator. (c) Mass change in water, facial sponge, Fe2O3/facial sponge, and Fe2O3@PPy solar evaporators under one sun illumination. (d) The long-term evaporation performance with minor disparity under one sun irradiation for continuous evaporation up to 8.3 h in seawater conditions. (e) Surface temperatures of Fe2O3@PPy solar evaporator and vapors. (f) The achieved evaporation rates and respective solar to vapor conversion efficiencies of the designed four evaporation systems.
Crystals 11 01509 g005
Figure 6. (a,b) Microscopical image of methyl orange/water solution and condensed water after evaporation through Fe2O3@PPy solar evaporator during Raman spectroscopy. (c) Raman spectra of methyl-orange dissolved water and condensed water show no peaks that show excellent degradation. (d,e) Concentration of the primary metals ions, and heavy metals ions before and after desalination. (f) Complete rejection of primary salt ions in stimulated seawater.
Figure 6. (a,b) Microscopical image of methyl orange/water solution and condensed water after evaporation through Fe2O3@PPy solar evaporator during Raman spectroscopy. (c) Raman spectra of methyl-orange dissolved water and condensed water show no peaks that show excellent degradation. (d,e) Concentration of the primary metals ions, and heavy metals ions before and after desalination. (f) Complete rejection of primary salt ions in stimulated seawater.
Crystals 11 01509 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, Y.; Arshad, N.; Irshad, M.S.; Ahmed, I.; Ahmad, S.; Alshahrani, L.A.; Yousaf, M.; Sayed, A.E.; Nauman, M. Fe2O3 Nanoparticles Deposited over Self-Floating Facial Sponge for Facile Interfacial Seawater Solar Desalination. Crystals 2021, 11, 1509. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121509

AMA Style

Lu Y, Arshad N, Irshad MS, Ahmed I, Ahmad S, Alshahrani LA, Yousaf M, Sayed AE, Nauman M. Fe2O3 Nanoparticles Deposited over Self-Floating Facial Sponge for Facile Interfacial Seawater Solar Desalination. Crystals. 2021; 11(12):1509. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121509

Chicago/Turabian Style

Lu, Yuzheng, Naila Arshad, Muhammad Sultan Irshad, Iftikhar Ahmed, Shafiq Ahmad, Lina Abdullah Alshahrani, Muhammad Yousaf, Abdelaty Edrees Sayed, and Muhammad Nauman. 2021. "Fe2O3 Nanoparticles Deposited over Self-Floating Facial Sponge for Facile Interfacial Seawater Solar Desalination" Crystals 11, no. 12: 1509. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11121509

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