More stories

  • in

    Urgent action is needed to restore the water sector in Ukraine

    Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.This is a summary of: Shumilova, O. et al. Impact of the Russia–Ukraine armed conflict on water resources and water infrastructure. Nat. Sustain. https://doi.org/10.1038/s41893-023-01068-x (2023). More

  • in

    Aquifer conditions, not irradiance determine the potential of photovoltaic energy for groundwater pumping across Africa

    PVWPS operationThe motor and the pump are built in together14 and the motor-pump set is submersed in the borehole under the water43. Control equipment is also installed between the PV modules and the motor-pump and/or integrated to the motor-pump set in the borehole14,17. This equipment allows the motor-pump to stop and also to operate the motor-pump and the PV modules at their best operating points14. Once the water is pumped, it might then be stored in a water tank to mitigate the variability of solar resources14,29. When pumping starts, a cone of depression of radius rc is formed and there is a drawdown Hb,d in the borehole (see Fig. 1). The higher the pumped flow rate, the higher the drawdown Hb,d and therefore the deeper the water in the borehole Hb. If Hb reaches the position of the motor-pump Hmp, the motor-pump automatically switches off, therefore preventing the motor-pump from running dry44. The motor-pump remains shut down during a period Δtshut, after which it makes an attempt to restart44.Input data processingWe observe in Table 1 that the datasets have varying spatial resolutions. In the article, we use the spatial resolution of the irradiance map, 0.2° (~22 km). Indeed, this resolution is sufficient for the purposes of this article and it allows to divide computing time and memory requirements by ~16 in comparison to the 0.05° resolution. At this 0.2° resolution, the total area of Africa of 30 million km2 is divided into 62,000 pixels. We apply this resolution of 0.2° to all datasets by nearest interpolation.No exact value of the static water depth Hb,s, transmissivity T and saturated thickness Hst are provided for each location by the original source but only a range of variation. For instance, for −15.8° (lat) & 21.9° (lon), the saturated thickness Hst is comprised between 25 and 100 m. In most cases, we consider the middle of the range (e.g., 62.5 m in the example). The only two exceptions are: when Hb,s is higher than 250 m, we consider 300 m (same for Hst); and, when Hb,s is between 0 and 7 m, we consider 7 m45. Due to the lack of available information, the input groundwater data provided in Table 1 are considered to remain constant over time.Reference46 provides complete irradiance data with a time step of 15 min from 2013 to 2020 across Africa. In this article, except mentioned otherwise, we use irradiance data from 2020 with a 30-min time step (by taking one point every two 15-min points), instead of all the available complete irradiance data. It divides computing time and memory requirements by ~16. Additionally, it produces reduced and acceptable deviations on the results. Indeed, for 100 randomly chosen locations, we simulated the pumped volume V, for the three considered PVWPS sizes, using (1) irradiance data from 2013 to 2020 with a 15-min time step and (2) irradiance data from 2020 with a 30-min time step. For these locations, the absolute error on volume V is systematically lower than 7.9% and the average absolute error is 2%. These results are coherent with the observed low influence of irradiance on the pumped volume in comparison to groundwater resources. Thanks to the consideration of this reduced irradiance vector, the random access memory (RAM) and the computing time required to obtain a map of final results (such as Fig. 5b) are respectively 38 Gb and 10 h (time for Intel Xeon E5-2643 3.3 GHz processors and 96 GB RAM, running on Debian 4.19.194-2), which is more reasonable.Atmospheric sub-modelFor each location, the irradiance on the plane of the PV modules Gpv at time t can be deduced from satellite data by47,48:$${G}_{{{{{{rm{pv}}}}}}}left(tright)={G}_{{{{{{rm{bn}}}}}}}left(tright){{cos }}left({{{{{rm{AOI}}}}}}left(t,theta ,alpha right)right)+{G}_{{{{{{rm{gh}}}}}}}left(tright)kappa frac{1-{{cos }}left(theta right)}{2}+{G}_{{{{{{rm{dh}}}}}}}left(tright)frac{1+{{cos }}left(theta right)}{2}$$
    (1)
    where κ is the albedo of the surrounding environment, θ and α are the tilt and azimuth of the PV modules and AOI is the angle of incidence between the sun’s rays and the PV modules. The albedo κ is taken equal to 0.2 because it corresponds to the albedo of cropland, which is a common environment in the rural areas considered49. In any case, additional simulations show that the value of the albedo has a negligible effect on the pumped volume V. AOI is computed using the MATLAB toolbox PVLIB developed by the Sandia National Laboratories50.For each location, the azimuth α and the tilt θ of the PV modules are chosen to maximize the irradiance on the plane of the PV modules Gpv. The azimuth α is taken equal to51:$$alpha =left{begin{array}{c}180^circ quad {{{{{rm{if}}}}}},phi , > ,0\ 0^circ quad {{{{{rm{if}}}}}},phi , < ,0end{array}right.$$ (2) where ϕ is the latitude of the location. The tilt is taken equal to51:$$theta =left{begin{array}{c}{{max }},(10,1.3793+(1.2011+(-0.014404+0.000080509phi )phi )phi )quad{{{{{rm{if}}}}}},phi > ,0\ {{min }},(-10,-0.41657+(1.4216+(0.024051+0.00021828phi )phi )phi )quad{{{{{rm{if}}}}}},phi , < ,0end{array}right.$$ (3) As evidenced by Eq. (3), the tilt should be higher than 10° or lower than −10°, so that the PV modules are tilted enough to be cleaned when it rains.Photovoltaic modules sub-modelConsidering that the maximum power point tracking of the PV modules is correctly performed, a simplified model to compute the power P produced by the modules is used:$$Pleft(tright)=frac{{G}_{{{{{{rm{pv}}}}}}}left(tright)}{{G}_{0}}{P}_{{{{{{rm{p}}}}}}}left(1-{c}_{{{{{{rm{pv}}}}}},{{{{{rm{loss}}}}}}}right)$$ (4) where G0 is the reference irradiance (1000 W m−2), Pp is the peak power of the PV modules in standard test conditions (STC) and cpv,loss is a coefficient that represents the losses (e.g., soiling, temperature, mismatch, wiring52,53) at the level of the PV modules. For the sake of simplicity, and as we consider a generic PVWPS, we consider that cpv,loss is independent of the operating point of the PV modules, of the time, and of the location. We take it constant, equal to a single value (see Table 2).Hydraulic sub-modelThe total dynamic head TDH between the motor-pump and the pipe output is given by54:$${TD}Hleft(tright)={H}_{{{{{{rm{b}}}}}}}(t)+{H}_{{{{{{rm{p}}}}}}}(t)$$ (5) where Hb is the water depth in the borehole and Hp is the additional head due to pressure losses in the pipe.The water depth in the borehole Hb is given by (see Fig. 1)42:$${H}_{{{{{{rm{b}}}}}}}(t)={H}_{{{{{{rm{b}}}}}},{{{{{rm{s}}}}}}}+{H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}(t)$$ (6) where Hb,s is the static water depth and Hb,d is the drawdown. The drawdown is composed of two parts:$${H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}(t)={H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}(t)+{H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}(t)$$ (7) where ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}(t)) is the head loss due to aquifer losses and ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}(t)) is the head loss due to borehole losses.The head loss due to aquifer losses ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}(t)) depends on the pumping flow rate Q, the aquifer transmissivity T, the borehole radius rb, and a length parameter rc representing the distance of water travel to replace the water pumped out. From dimensional analysis, we expect that (tfrac{{H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}left(tright)cdot T}{Qleft(tright)}) should be a function of (tfrac{{r}_{{{{{{rm{c}}}}}}}}{{r}_{{{{{{rm{b}}}}}}}}). We thus propose the following model for ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}), which is derived from Thiem equation55:$${H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}left(tright)=frac{{{{{{rm{ln}}}}}}left(frac{{r}_{{{{{{rm{c}}}}}}}}{{r}_{{{{{{rm{b}}}}}}}}right)}{2pi T}Qleft(tright)$$ (8) where rc can be considered the effective radius of the cone of depression. This model satisfies horizontal, radial and steady Darcy flow in a uniform, homogeneous and isotropic aquifer. It captures the essential features for aquifer losses: ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{a}}}}}}}(t)) proportional to pumped flow rate and inversely proportional to transmissivity45. Though the flow is transient, only simplified steady-state models, as the one of Eq. (8), can be applied with the available information as dynamic models would require pumping tests. Furthermore, we consider that the radius of the cone of depression rc is comprised between 100 and 1000 m and, to correlate it to a measured quantity, that it depends linearly on the groundwater recharge R: for the lowest recharge (0 m/year), rc is equal to 1000 m; for the highest one (0.2947 m year−1), rc is equal to 100 m; in-between, rc is obtained linearly from the recharge (rc = 1000–3054 · R). Thus, groundwater recharge R is used to constrain the size of the cone of depression.The head loss due to borehole losses ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}(t)) is given by56:$${H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}left(tright)=beta {Q}{left(tright)}^{2}$$ (9) where β is a coefficient related to the borehole design. For the yields considered in this article, ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}(t)) usually remains lower than a few meters but, as ({H}_{{{{{{rm{b}}}}}},{{{{{rm{d}}}}}}}^{{{{{{rm{b}}}}}}}(t)) depends on the square of the pumped flow rate, it may be more important for larger abstraction capacities.The additional head due to pipe losses Hp is given by57:$${H}_{{{{{{rm{p}}}}}}}left(tright)={H}_{{{{{{rm{p}}}}}},{{{{{rm{ma}}}}}}}left(tright)+{H}_{{{{{{rm{p}}}}}},{{{{{rm{mi}}}}}}}left(tright)$$ (10) where Hp,ma(t) corresponds to losses that occur along the pipe length (also called “major losses”) and Hp,mi(t) corresponds to losses at junctions such as elbows and curvatures (also called “minor losses”). Hp,ma(t) is given by57:$${H}_{{{{{{rm{p}}}}}},{{{{{rm{ma}}}}}}}left(tright)=frac{8f}{{pi }^{2}g{D}_{{{{{{rm{p}}}}}}}^{5}}{L}_{{{{{{rm{p}}}}}}}Q{left(tright)}^{2}$$ (11) where g is the gravitational acceleration (9.81 m s−2), Dp is the pipe diameter, Lp is the pipe length, Q is the pumped flow rate, and f is the friction coefficient between the water and the pipe. We approximate the pipe length Lp to be equal to the depth of the motor-pump Hmp (see Fig. 1). The expression of f depends on the value of the Reynolds number ({{{{{rm{Re}}}}}}=tfrac{4Q}{pi {D}_{{{{{{rm{p}}}}}}}w}), where w is the water kinematic viscosity (taken equal to 1 × 10−6 m2 s−1)57: for Re More

  • in

    Scalable and switchable CO2-responsive membranes with high wettability for separation of various oil/water systems

    Dai, H., Dong, Z. & Jiang, L. Directional liquid dynamics of interfaces with superwettability. Sci. Adv. 6, eabb5528 (2020).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Chaudhury, M. K. & Whitesides, G. M. How to make water run uphill. Science 256, 1539–1541 (1992).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Kwon, G. et al. Visible light guided manipulation of liquid wettability on photoresponsive surfaces. Nat. Commun. 8, 14968 (2017).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Liu, M., Wang, S. & Jiang, L. Nature-inspired superwettability systems. Nat. Rev. Mater. 2, 17036 (2017).Article 
    ADS 
    CAS 

    Google Scholar 
    Chu, K. H., Xiao, R. & Wang, E. N. Uni-directional liquid spreading on asymmetric nanostructured surfaces. Nat. Mater. 9, 413–417 (2010).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Kota, A. K., Kwon, G., Choi, W., Mabry, J. M. & Tuteja, A. Hygro-responsive membranes for effective oil–water separation. Nat. Commun. 3, 1025 (2012).Article 
    ADS 
    PubMed 

    Google Scholar 
    Zhang, G. et al. Processing supramolecular framework for free interconvertible liquid separation. Nat. Commun. 11, 425 (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, B., Liang, W., Guo, Z. & Liu, W. Biomimetic super-lyophobic and super-lyophilic materials applied for oil/water separation: a new strategy beyond nature. Chem. Soc. Rev. 44, 336–361 (2015).Article 
    PubMed 

    Google Scholar 
    Jiang, Y. et al. Crystal structural and mechanism of a calcium-gated potassium channel. Nature 417, 523–526 (2002).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Liu, Z., Wang, W., Xie, R., Ju, X. J. & Chu, L. Y. Stimuli-responsive smart gating membranes. Chem. Soc. Rev. 45, 460–475 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hou, X. Smart gating multi-scale pore/channel-based membranes. Adv. Mater. 28, 7049–7064 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Qu, R. et al. Aminoazobenzene@Ag modified meshes with large extent photo-response: towards reversible oil/water removal from oil/water mixtures. Chem. Sci. 10, 4089–4096 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Qu, R. et al. Photothermally induced in situ double emulsion separation by a carbon nanotube/poly(Nisopropylacrylamide) modified membrane with superwetting properties. J. Mater. Chem. A 8, 7677–7686 (2020).Article 
    CAS 

    Google Scholar 
    Hu, L. et al. Photothermal-responsive single-walled carbon nanotube-based ultrathin membranes for on/off switchable separation of oil-in-water nanoemulsions. ACS Nano 9, 4835–4842 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Cheng, B. et al. Development of smart poly(vinylidene fluoride)-graft-poly(acrylic acid) tree-like nanofiber membrane for pH-responsive oil/water separation. J. Membr. Sci. 534, 1–8 (2017).Article 
    CAS 

    Google Scholar 
    Zhang, L., Zhang, Z. & Wang, P. Smart surfaces with switchable superoleophilicity and superoleophobicity in aqueous media: toward controllable oil/water separation. NPG Asia Mater. 4, e8 (2012).Article 

    Google Scholar 
    Li, J. J., Zhou, Y. N. & Luo, Z. H. Smart fiber membrane for pH-induced oil/water separation. ACS Appl. Mater. Inter. 7, 19643–19650 (2015).Article 
    CAS 

    Google Scholar 
    Kwon, G. et al. On-demand separation of oil-water mixtures. Adv. Mater. 24, 3666 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Liu, Y., Zhao, L., Lin, J. & Yang, S. Electrodeposited surfaces with reversibly switching interfacial properties. Sci. Adv. 5, eaax0380 (2019).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, W. et al. Thermo-driven controllable emulsion separation by a polymer-decorated membrane with switchable wettability. Angew. Chem. Int. Ed. 57, 5740 (2018).Article 
    CAS 

    Google Scholar 
    Ou, R., Wei, J., Jiang, L., Simon, G. P. & Wang, H. Robust thermos-responsive polymer composite membrane with switchable superhydrophilicity and superhydrophobicity for efficient oil-water separation. Environ. Sci. Technol. 50, 906–914 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Rana, D. & Matsuura, T. Surface modifications for antifouling membranes. Chem. Rev. 110, 2448–2471 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Dutta, K. & De, S. Smart responsive materials for water purification: an overview. J. Mater. Chem. A 5, 22095–22112 (2017).Article 
    CAS 

    Google Scholar 
    Dong, L. & Zhao, Y. CO2-switchable membranes: structures, functions, and separation applications in aqueous medium. J. Mater. Chem. A 8, 16738–16746 (2020).Article 
    CAS 

    Google Scholar 
    Cunningham, M. F. & Jessop, P. G. Carbon dioxide-switchable polymers: where are the future opportunities? Macromolecules 52, 6801–6816 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Zhang, Q., Lei, L. & Zhu, S. Gas-responsive polymers. ACS Macro Lett. 6, 515–522 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Liu, H., Lin, S., Feng, Y. & Theato, P. CO2-responsive polymer materials. Polym. Chem. 8, 12–23 (2017).Article 

    Google Scholar 
    Lin, S. J., Shang, J. J. & Theato, P. Facile fabrication of CO2‑responsive nanofibers from photo-cross-linked poly(pentafluorophenyl acrylate) nanofibers. ACS Macro Lett. 7, 431–436 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lin, S. J. & Theato, P. CO2-responsive polymers. Macromol. Rapid Commun. 34, 1118–1133 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yan, Q. & Zhao, Y. Block copolymer self-assembly controlled by the “green” gas stimulus of carbon dioxide. Chem. Commun. 50, 11631–11641 (2014).Article 
    CAS 

    Google Scholar 
    Yan, Q. & Zhao, Y. CO2‑stimulated diversiform deformations of polymer assemblies. J. Am. Chem. Soc. 135, 16300–16303 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Darabi, A., Jessop, P. G. & Cunningham, M. F. CO2-responsive polymeric materials: synthesis, self-assembly, and functional applications. Chem. Soc. Rev. 45, 4391–4436 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Che, H. et al. CO2-responsive nanofibrous membranes with switchable oil/water wettability. Angew. Chem. Int. Ed. 54, 8934–8938 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Lei, L., Zhang, Q., Shi, S. & Zhu, S. Highly porous poly(high internal phase emulsion) membranes with “open-cell” structure and CO2‑switchable wettability used for controlled oil/water separation. Langmuir 33, 11936–11944 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Mo, J., Sha, J., Li, D., Li, Z. & Chen, Y. Fluid release pressure for nanochannels: the Young-Laplace equation using the effective contact angle. Nanoscale 11, 8408–8415 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Seveno, D., Blake, T. D. & Coninck, J. Young’s equation at the nanoscale. Phys. Rev. Lett. 111, 096101 (2013).Article 
    ADS 
    PubMed 

    Google Scholar 
    Yang, Z., He, C., Suia, H., He, L. & Li, X. Recent advances of CO2-responsive materials in separations. J. CO2 Util. 30, 79–99 (2019).Article 
    CAS 

    Google Scholar 
    Tauhardt, L. et al. Zwitterionic poly(2-oxazoline)s as promising candidates for blood contacting applications. Polym. Chem. 5, 5751 (2014).Article 
    CAS 

    Google Scholar 
    Zhang, X. et al. Janus poly(vinylidene fluoride)-graft-(TiO2 nanoparticles and PFDS) membranes with loose architecture and asymmetric wettability for efficient switchable separation of surfactant-stabilized oil/water emulsions. J. Membr. Sci. 640, 119837 (2021).Article 
    CAS 

    Google Scholar 
    Feng, X. & Jiang, L. Design and creation of superwetting/antiwetting surfaces. Adv. Mater. 18, 3063–3078 (2006).Article 
    CAS 

    Google Scholar 
    Tao, M., Xue, L., Liu, F. & Jiang, L. An intelligent superwetting PVDF membrane showing switchable transport performance for oil/water separation. Adv. Mater. 26, 2943–2948 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Li, X. et al. Universal and tunable liquid-liquid separation by nanoparticle-embedded gating membranes based on a self-defined interfacial parameter. Nat. Commun. 12, 80 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Peng, B. et al. Cellulose-based materials in wastewater treatment of petroleum industry. Green. Energy Environ. 5, 37–49 (2020).Article 

    Google Scholar 
    Huang, K. et al. Cation-controlled wetting properties of vermiculite membranes and its promise for fouling resistant oil-water separation. Nat. Commun. 11, 1097 (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Haase, M. F. et al. Multifunctional nanocomposite hollow fiber membranes by solvent transfer induced phase separation. Nat. Commun. 8, 1234 (2017).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    J. Zhang, L. Liu, Y. Si, J. Yu, B. Ding. Electrospun nanofibrous membranes: an effective arsenal for the purification of emulsified oily wastewater. Adv. Funct. Mater. 30, 2002192 (2020).Kim, S., Kim, K., Jun, G. & Hwang, W. Wood-nanotechnology-based membrane for the efficient purification of oil-in-water emulsions. ACS Nano 14, 17233–17240 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kralchevsky, P. A. & Nagayama, K. Capillary interactions between particles bound to interfaces, liquid films and biomembranes. Adv. Colloid Interfacace Sci. 85, 145–192 (2000).Article 
    CAS 

    Google Scholar 
    Dehkordi, T. F., Shirin-Abadi, A. R., Karimipour, K. & Mahdavian, A. R. CO2-, electric potential-, and photo-switchable-hydrophilicity membrane (x-SHM) as an efficient color-changeable tool for oil/water separation. Polymer 212, 123250 (2021).Article 
    CAS 

    Google Scholar 
    Huang, X., Mutlu, H. & Theato, P. A CO2-gated anodic aluminum oxide based nanocomposite membrane for de-emulsification. Nanoscale 12, 21316–21324 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Shirin-Abadi, A. R., Gorji, M., Rezaee, S., Jessop, P. G. & Cunningham, M. F. CO2-Switchable-hydrophilicity membrane (CO2-SHM) triggered by electric potential: faster switching time along with efficient oil/water separation. Chem. Commun. 54, 8478–8481 (2018).Article 

    Google Scholar 
    Abraham, S., Kumaranab, S. K. & Montemagno, C. D. Gas-switchable carbon nanotube/polymer hybrid membrane for separation of oil-in-water emulsions. RSC Adv. 7, 39465–39470 (2017).Article 
    ADS 
    CAS 

    Google Scholar 
    Liu, Y. et al. Separate reclamation of oil and surfactant from oil-in-water emulsion with a CO2-responsive material. Environ. Sci. Technol. 56, 9651–9660 (2022).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Zhang, Q. et al. CO2‑Switchable membranes prepared by immobilization of CO2‑breathing microgels. ACS Appl. Mater. Inter. 9, 44146–44151 (2017).Article 
    CAS 

    Google Scholar 
    Kung, C. H., Zahiri, B., Sow, P. K. & Mérida, W. On-demand oil-water separation via low-voltage wettability switching of core-shell structures on copper substrates. Appl. Surf. Sci. 444, 15–27 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Du, L., Quan, X., Fan, X., Chen, S. & Yu, H. Electro-responsive carbon membranes with reversible superhydrophobicity/superhydrophilicity switch for efficient oil/water separation. Sep. Purif. Technol. 210, 891–899 (2019).Article 
    CAS 

    Google Scholar 
    Wu, J. et al. A 3D smart wood membrane with high flux and efficiency for separation of stabilized oil/water emulsions. J. Hazard. Mater. 441, 129900 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yang, C. et al. Facile fabrication of durable mesh with reversible photo-responsive wettability for smart oil/water separation. Prog. Org. Coat. 160, 106520 (2021).Article 
    CAS 

    Google Scholar 
    Chen, Z. et al. Smart light responsive polypropylene membrane switching reversibly between hydrophobicity and hydrophilicity for oily water separation. J. Membr. Sci. 638, 119704 (2021).Article 
    CAS 

    Google Scholar 
    Li, J. J., Zhu, L. T. & Luo, Z. H. Electrospun fibrous membrane with enhanced swithchable oil/water wettability for oily water separation. Chem. Eng. J. 287, 474–481 (2016).Article 
    CAS 

    Google Scholar 
    Wang, Y. et al. Temperature-responsive nanofibers for controllable oil/water separation. RSC Adv. 5, 51078–51085 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Ding, Y. et al. One-step fabrication of a micro/nanosphere-coordinated dual stimulus-responsive nanofibrous membrane for intelligent antifouling and ultrahigh permeability of viscous water-in-oil emulsions. ACS Appl. Mater. Inter. 13, 27635–27644 (2021).Article 
    CAS 

    Google Scholar 
    Xiang, Y., Shen, J., Wang, Y., Liu, F. & Xue, L. A pH-responsive PVDF membrane with superwetting properties for the separation of oil and water. RSC Adv. 5, 23530–23539 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Dou, Y. L. et al. Dual-responsive polyacrylonitrile-based electrospun membrane for controllable oil-water separation. J. Hazard. Mater. 438, 129565 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Li, J. J., Zhou, Y. N. & Luo, Z. H. Mussel-inspired V-shaped copolymer coating for intelligent oil/water separation. Chem. Eng. J. 322, 693–701 (2017).Article 
    CAS 

    Google Scholar 
    Zhou, Y. N., Li, J. J. & Lu, Z. H. Toward efficient water/oil separation material: effect of copolymer composition on pH-responsive wettability and separation performance. AIChE J. 62, 1758–1770 (2016).Article 
    CAS 

    Google Scholar 
    Xu, Z. et al. Fluorine-free superhydrophobic coatings with pH-induced wettability transition for controllable oil−water separation. ACS Appl. Mater. Inter. 8, 5661–5667 (2016).Article 
    CAS 

    Google Scholar 
    Dang, Z., Liu, L., Li, Y., Xiang, Y. & Guo, G. In situ and ex situ pH-responsive coatings with switchable wettability for controllable oil/water separation. ACS Appl. Mater. Inter. 8, 31281–31288 (2016).Article 
    CAS 

    Google Scholar 
    Cai, Y. et al. A smart membrane with antifouling capability and switchable oil wettability for high-efficiency oil/water emulsions separation. J. Membr. Sci. 555, 69–77 (2018).Article 
    CAS 

    Google Scholar 
    Zhang, Z. et al. A reusable, biomass-derived, and pH-responsive collagen fiber based oil absorbent material for effective separation of oil-in-water emulsions. Colloid Surf. A 633, 127906 (2022).Article 
    CAS 

    Google Scholar 
    Zhang, X., Liu, C., Yang, J., Huang, X. J. & Xu, Z. K. Wettability switchable membranes for separating both oil-in-water and water-in-oil emulsions. J. Membr. Sci. 624, 118976 (2021).Article 
    CAS 

    Google Scholar 
    Yang, W., Li, J., Zhou, P., Zhu, L. & Tang, H. Superhydrophobic copper coating: Switchable wettability, on-demand oil-water separation, and antifouling. Chem. Eng. J. 327, 849–854 (2017).Article 
    CAS 

    Google Scholar 
    Zhai, H. et al. Crown ether modified membranes for Na+-responsive controllable emulsion separation suitable for hypersaline environments. J. Mater. Chem. A 8, 2684–2690 (2020).Article 
    CAS 

    Google Scholar  More

  • in

    Oil-in-water nanoemulsions for better nanofiltration membranes

    Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain
    the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in
    Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles
    and JavaScript. More

  • in

    Environmental impact of direct lithium extraction from brines

    Trahey, L. et al. Energy storage emerging: a perspective from the Joint Center for Energy Storage Research. Proc. Natl Acad. Sci. USA 117, 12550–12557 (2020).Article 

    Google Scholar 
    Koohi-Fayegh, S. & Rosen, M. A. A review of energy storage types, applications and recent developments. J. Energy Storage 27, 101047 (2020).Article 

    Google Scholar 
    Dehghani-Sanij, A. R., Tharumalingam, E., Dusseault, M. B. & Fraser, R. Study of energy storage systems and environmental challenges of batteries. Renew. Sustain. Energy Rev. 104, 192–208 (2019).Article 

    Google Scholar 
    Olivetti, E. A., Ceder, G., Gaustad, G. G. & Fu, X. Lithium-ion battery supply chain considerations: analysis of potential bottlenecks in critical metals. Joule 1, 229–243 (2017).Article 

    Google Scholar 
    Global EV Outlook (IEA, 2021); https://www.iea.org/reports/global-ev-outlook-2021.Tabelin, C. B. et al. Towards a low-carbon society: a review of lithium resource availability, challenges and innovations in mining, extraction and recycling, and future perspectives. Miner. Eng. 163, 106743 (2021).Article 

    Google Scholar 
    The role of Critical World Energy Outlook Special Report Minerals in Clean Energy Transitions (IEA, 2022); https://iea.blob.core.windows.net/assets/ffd2a83b-8c30-4e9d-980a-52b6d9a86fdc/TheRoleofCriticalMineralsinCleanEnergyTransitions.pdf.Xu, C. et al. Future material demand for automotive lithium-based batteries. Commun. Mater. 1, 99 (2020).Article 

    Google Scholar 
    Mineral Commodity Summaries. LITHIUM (USGS, 2021); https://pubs.usgs.gov/periodicals/mcs2021/mcs2021-lithium.pdf.Kesler, S. E. et al. Global lithium resources: relative importance of pegmatite, brine and other deposits. Ore Geol. Rev. 48, 55–69 (2012).Article 

    Google Scholar 
    Alessia, A., Alessandro, B., Maria, V. G., Carlos, V. A. & Francesca, B. Challenges for sustainable lithium supply: a critical review. J. Clean. Prod. 300, 126954 (2021).Article 

    Google Scholar 
    Tadesse, B., Makuei, F., Albijanic, B. & Dyer, L. The beneficiation of lithium minerals from hard rock ores: a review. Miner. Eng. 131, 170–184 (2019).Article 

    Google Scholar 
    Vikström, H., Davidsson, S. & Höök, M. Lithium availability and future production outlooks. Appl. Energy 110, 252–266 (2013).Article 

    Google Scholar 
    Sanjuan, B. et al. Major geochemical characteristics of geothermal brines from the Upper Rhine Graben granitic basement with constraints on temperature and circulation. Chem. Geol. 428, 27–47 (2016).Article 

    Google Scholar 
    Sanjuan, B. et al. Lithium-rich geothermal brines in Europe: an up-date about geochemical characteristics and implications for potential Li resources. Geothermics 101, 102385 (2022).Article 

    Google Scholar 
    Stringfellow, W. T. & Dobson, P. F. Technology for the recovery of lithium from geothermal brines. Energies 14, 6805 (2021).Article 

    Google Scholar 
    Dugamin, E. J. M. et al. Groundwater in sedimentary basins as potential lithium resource: a global prospective study. Sci. Rep. 11, 21091 (2021).Article 

    Google Scholar 
    Flexer, V., Baspineiro, C. F. & Galli, C. I. Lithium recovery from brines: a vital raw material for green energies with a potential environmental impact in its mining and processing. Sci. Total Environ. 639,, 1188–1204 (2018).Article 

    Google Scholar 
    Garrett, D. E. Handbook of Lithium and Natural Calcium Chloride https://doi.org/10.1016/B978-0-12-276152-2.X5035-X (2004).Article 

    Google Scholar 
    Mudd, G. M. Sustainable/responsible mining and ethical issues related to the Sustainable Development Goals. Geol. Soc. London, Spec. Publ. 508, 187 LP–199 (2021).Article 

    Google Scholar 
    Jerez, B., Garcés, I. & Torres, R. Lithium extractivism and water injustices in the Salar de Atacama, Chile: the colonial shadow of green electromobility. Polit. Geogr. 87, 102382 (2021).Article 

    Google Scholar 
    Alam, M. A. & Sepúlveda, R. Environmental degradation through mining for energy resources: the case of the shrinking Laguna Santa Rosa wetland in the Atacama Region of Chile. Energy Geosci. 3, 182–190 (2022).Article 

    Google Scholar 
    Hailes, O. Lithium in international law: trade, investment, and the pursuit of supply chain justice. J. Int. Econ. Law 25, 148–170 (2022).Article 

    Google Scholar 
    Bustos-Gallardo, B., Bridge, G. & Prieto, M. Harvesting lithium: water, brine and the industrial dynamics of production in the Salar de Atacama. Geoforum 119, 177–189 (2021).Article 

    Google Scholar 
    Agusdinata, D. B., Liu, W., Eakin, H. & Romero, H. Socio-environmental impacts of lithium mineral extraction: towards a research agenda. Environ. Res. Lett. 13, 123001 (2018).Article 

    Google Scholar 
    Liu, W. & Agusdinata, D. B. Interdependencies of lithium mining and communities sustainability in Salar de Atacama, Chile. J. Clean. Prod. 260, 120838 (2020).Article 

    Google Scholar 
    Ellingsen, L. A. W., Singh, B. & Strømman, A. H. The size and range effect: lifecycle greenhouse gas emissions of electric vehicles. Environ. Res. Lett. 11, 054010 (2016).Article 

    Google Scholar 
    Ambrose, H. & Kendall, A. Understanding the future of lithium: part 2, temporally and spatially resolved life-cycle assessment modeling. J. Ind. Ecol. 24, 90–100 (2020).Article 

    Google Scholar 
    Porzio, J. & Scown, C. D. Life-cycle assessment considerations for batteries and battery materials. Adv. Energy Mater. 11, 2100771 (2021).Article 

    Google Scholar 
    Pell, R. et al. Towards sustainable extraction of technology materials through integrated approaches. Nat. Rev. Earth Environ. 2, 665–679 (2021).Article 

    Google Scholar 
    Stamp, A., Lang, D. J. & Wäger, P. A. Environmental impacts of a transition toward e-mobility: the present and future role of lithium carbonate production. J. Clean. Prod. 23, 104–112 (2012).Article 

    Google Scholar 
    Ejeian, M., Grant, A., Shon, H. K. & Razmjou, A. Is lithium brine water? Desalination 518, 115169 (2021). Discussion of why brine water should be considered in environmental assessments.Article 

    Google Scholar 
    Marazuela, M. A., Vázquez-Suñé, E., Ayora, C. & García-Gil, A. Towards more sustainable brine extraction in salt flats: learning from the Salar de Atacama. Sci. Total. Environ. 703, 135605 (2020). Conceptual hydrogeological modelling, calibrated with field data, proposing clever strategies to minimize water impacts during brine pumping.Article 

    Google Scholar 
    Marazuela, M. A., Vázquez-Suñé, E., Ayora, C., García-Gil, A. & Palma, T. Hydrodynamics of salt flat basins: the Salar de Atacama example. Sci. Total. Environ. 651, 668–683 (2019).Article 

    Google Scholar 
    Marazuela, M. A., Vázquez-Suñé, E., Ayora, C., García-Gil, A. & Palma, T. The effect of brine pumping on the natural hydrodynamics of the Salar de Atacama: the damping capacity of salt flats. Sci. Total. Environ. 654, 1118–1131 (2019).Article 

    Google Scholar 
    Houston, J., Butcher, A., Ehren, P., Evans, K. & Godfrey, L. The evaluation of brine prospects and the requirement for modifications to filing standards. Econ. Geol. 106, 1125–1239 (2011). Conceptual hydrogeological modelling about brine pumping and freshwater recharge.Article 

    Google Scholar 
    Liu, W., Agusdinata, D. B. & Myint, S. W. Spatiotemporal patterns of lithium mining and environmental degradation in the Atacama Salt Flat, Chile. Int. J. Appl. Earth Obs. Geoinf. 80, 145–156 (2019). Pioneering publication with solid data highlighting the environmental impacts related to lithium mining from continental brines.
    Google Scholar 
    Gutierrez, J. S. et al. Climate change and lithium mining influence flamingo abundance in the Lithium Triangle. Proc. R. Soc. B Biol. Sci. 289, 20212388 (2022).Article 

    Google Scholar 
    Marazuela, M. A. et al. 3D mapping, hydrodynamics and modelling of the freshwater-brine mixing zone in salt flats similar to the Salar de Atacama (Chile). J. Hydrol. 561, 223–235 (2018).Article 

    Google Scholar 
    Rosen, M. R. The importance of groundwater in playas: a review of playa classifications and the sedimentology and hydrology of playas. in Paleoclimate and Basin Evolution of Playa Systems Vol. 289 (ed. Rosen, M. R.) (Geological Society of America, 1994).Currey, D. R. & Sack, D. Hemiarid Lake Basins: Hydrographic Patterns BT — Geomorphology of Desert Environments (eds Parsons, A. J. & Abrahams, A. D.) 471–487 (Springer Netherlands, 2009). https://doi.org/10.1007/978-1-4020-5719-9_15.Marconi, P., Arengo, F. & Clark, A. The arid Andean plateau waterscapes and the Lithium Triangle: flamingos as flagships for conservation of high-altitude wetlands under pressure from mining development. Wetl. Ecol. Manag. https://doi.org/10.1007/s11273-022-09872-6 (2022).Article 

    Google Scholar 
    Gajardo, G. & Redón, S. Andean hypersaline lakes in the Atacama Desert, northern Chile: between lithium exploitation and unique biodiversity conservation. Conserv. Sci. Pract. 1, e94 (2019).
    Google Scholar 
    Boualleg, M. & Burdet, F. A. P. Method of preparing an adsorbent material shaped in the absence of binder and method of extracting lithium from saline solutions using said material. US patent WO/097202 Al. (2015).Chen, S., Zhang, Q., Andrews-Speed, P. & Mclellan, B. Quantitative assessment of the environmental risks of geothermal energy: a review. J. Environ. Manage. 276, 111287 (2020).Article 

    Google Scholar 
    Megalooikonomou, K. G., Parolai, S. & Pittore, M. Toward performance-driven seismic risk monitoring for geothermal platforms: development of ad hoc fragility curves. Geotherm. Energy 6, 8 (2018).Article 

    Google Scholar 
    Bosia, C., Mouchot, J., Ravier, G., Seibel, O. & Genter, A. Evolution of brine geochemical composition during operation of EGS geothermal plants (Alsace, France). In Proc. 46th Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, California, February 15–17, 2021 SGP-TR-218 (2021).Tyszer, M., Tomaszewska, B. & Kabay, N. Desalination of geothermal wastewaters by membrane processes: strategies for environmentally friendly use of retentate streams. Desalination 520, 115330 (2021).Article 

    Google Scholar 
    Chagnes, A. & Światowska, J. Lithium Process Chemistry: Resources, Extraction, Batteries and Recycling (Elsevier, 2015).Khalil, A., Mohammed, S., Hashaikeh, R. & Hilal, N. Lithium recovery from brine: recent developments and challenges. Desalination 528, 115611 (2022).Article 

    Google Scholar 
    Meng, Z. et al. Highly flexible interconnected Li+ ion-sieve porous hydrogels with self-regulating nanonetwork structure for marine lithium recovery. Chem. Eng. J. 445, 136780 (2022).Article 

    Google Scholar 
    Marthi, R. & Smith, Y. R. Application and limitations of a H2TiO3 — diatomaceous earth composite synthesized from titania slag as a selective lithium adsorbent. Sep. Purif. Technol. 254, 117580 (2021).Article 

    Google Scholar 
    Li, X. et al. Amorphous TiO2-derived large-capacity lithium ion sieve for lithium recovery. Chem. Eng. Technol. 43, 1784–1791 (2020).Article 

    Google Scholar 
    Zhou, Z. et al. Recovery of lithium from salt-lake brines using solvent extraction with TBP as extractant and FeCl3 as co-extraction agent. Hydrometallurgy 191, 105244 (2020).Article 

    Google Scholar 
    Song, J., Huang, T., Qiu, H., Li, X. M. & He, T. Recovery of lithium from salt lake brine of high Mg/Li ratio using Na[FeCl4 ∗ 2TBP] as extractant: thermodynamics, kinetics and processes. Hydrometallurgy 173, 63–70 (2017).Article 

    Google Scholar 
    Li, R. et al. Novel ionic liquid as co-extractant for selective extraction of lithium ions from salt lake brines with high Mg/Li ratio. Sep. Purif. Technol. 277, 119471 (2021).Article 

    Google Scholar 
    Shi, C. et al. Solvent extraction of lithium from aqueous solution using non-fluorinated functionalized ionic liquids as extraction agents. Sep. Purif. Technol. 172, 473–479 (2017).Article 

    Google Scholar 
    Gmar, S. & Chagnes, A. Recent advances on electrodialysis for the recovery of lithium from primary and secondary resources. Hydrometallurgy 189, 105124 (2019).Article 

    Google Scholar 
    Li, X. et al. Membrane-based technologies for lithium recovery from water lithium resources: a review. J. Memb. Sci. 591, 117317 (2019).Article 

    Google Scholar 
    Zhao, Z., Liu, G., Jia, H. & He, L. Sandwiched liquid-membrane electrodialysis: lithium selective recovery from salt lake brines with high Mg/Li ratio. J. Memb. Sci. 596, 117685 (2020).Article 

    Google Scholar 
    Li, Q. et al. Efficiently rejecting and concentrating Li+ by nanofiltration membrane under a reversed electric field. Desalination 535, 115825 (2022).Article 

    Google Scholar 
    Li, Y., Zhao, Y. J., Wang, H. & Wang, M. The application of nanofiltration membrane for recovering lithium from salt lake brine. Desalination 468, 114081 (2019).Article 

    Google Scholar 
    He, R. et al. Unprecedented Mg2+/Li+ separation using layer-by-layer based nanofiltration hollow fiber membranes. Desalination 525, 115492 (2022).Article 

    Google Scholar 
    Calvo, E. J. Electrochemical methods for sustainable recovery of lithium from natural brines and battery recycling. Curr. Opin. Electrochem. 15, 102–108 (2019).Article 

    Google Scholar 
    Battistel, A., Palagonia, M. S., Brogioli, D., La Mantia, F. & Trócoli, R. Electrochemical methods for lithium recovery: a comprehensive and critical review. Adv. Mater. 32, 1905440 (2020). Critical review about electrochemical ion pumping technologies, with suggestions about which experimental parameters need to be assessed. Not all electrochemical technologies are reviewed.Article 

    Google Scholar 
    Calvo, E. J. Direct lithium recovery from aqueous electrolytes with electrochemical ion pumping and lithium intercalation. ACS Omega 6, 35213–35220 (2021).Article 

    Google Scholar 
    He, L. et al. New insights into the application of lithium-ion battery materials: selective extraction of lithium from brines via a rocking-chair lithium-ion battery system. Glob. Chall. 2, 1700079 (2018).Article 

    Google Scholar 
    Liu, D., Xu, W., Xiong, J., He, L. & Zhao, Z. Electrochemical system with LiMn2O4 porous electrode for lithium recovery and its kinetics. Sep. Purif. Technol. 270, 118809 (2021).Article 

    Google Scholar 
    Liu, D., Zhao, Z., Xu, W., Xiong, J. & He, L. A closed-loop process for selective lithium recovery from brines via electrochemical and precipitation. Desalination 519, 115302 (2021).Article 

    Google Scholar 
    Liu, D., Li, Z., He, L. & Zhao, Z. Facet engineered Li3PO4 for lithium recovery from brines. Desalination 514, 115186 (2021).Article 

    Google Scholar 
    Lai, X., Xiong, P. & Zhong, H. Extraction of lithium from brines with high Mg/Li ratio by the crystallization–precipitation method. Hydrometallurgy 192, 105252 (2020).Article 

    Google Scholar 
    Mendieta–George, D., Pérez–Garibay, R., Solís–Rodríguez, R., Fuentes–Aceituno, J. C. & Alvarado–Gómez, A. Study of the direct production of lithium phosphate with pure synthetic solutions and membrane electrolysis. Miner. Eng. 185, 107713 (2022).Article 

    Google Scholar 
    Cerda, A. et al. Recovering water from lithium-rich brines by a fractionation process based on membrane distillation–crystallization. J. Water Process. Eng 41, 102063 (2021). Membrane distillation-based DLE proposal to recover freshwater during brine concentration from high-salinity brines.Article 

    Google Scholar 
    Quist-Jensen, C. A., Ali, A., Mondal, S., Macedonio, F. & Drioli, E. A study of membrane distillation and crystallization for lithium recovery from high-concentrated aqueous solutions. J. Memb. Sci. 505, 167–173 (2016).Article 

    Google Scholar 
    Zhou, G. et al. Progress in electrochemical lithium ion pumping for lithium recovery. J. Energy Chem. 59, 431–445 (2021).Article 

    Google Scholar 
    Xu, P. et al. Materials for lithium recovery from salt lake brine. J. Mater. Sci. 56, 16–63 (2021).Article 

    Google Scholar 
    Wang, J. et al. Electrochemical technologies for lithium recovery from liquid resources: a review. Renew. Sustain. Energy Rev. 154, 111813 (2022).Article 

    Google Scholar 
    Sun, Y., Wang, Q., Wang, Y., Yun, R. & Xiang, X. Recent advances in magnesium/lithium separation and lithium extraction technologies from salt lake brine. Sep. Purif. Technol. 256, 117807 (2021).Article 

    Google Scholar 
    Nie, X.-Y., Sun, S.-Y., Sun, Z., Song, X. & Yu, J.-G. Ion-fractionation of lithium ions from magnesium ions by electrodialysis using monovalent selective ion-exchange membranes. Desalination 403, 128–135 (2017).Article 

    Google Scholar 
    Ding, D., Yaroshchuk, A. & Bruening, M. L. Electrodialysis through nafion membranes coated with polyelectrolyte multilayers yields >99% pure monovalent ions at high recoveries. J. Memb. Sci. 647, 120294 (2022).Article 

    Google Scholar 
    Li, Q. et al. Ultrahigh-efficient separation of Mg2+/Li+ using an in-situ reconstructed positively charged nanofiltration membrane under an electric field. J. Memb. Sci. 641, 119880 (2022).Article 

    Google Scholar 
    Qiu, Y. et al. Integration of selectrodialysis and selectrodialysis with bipolar membrane to salt lake treatment for the production of lithium hydroxide. Desalination 465, 1–12 (2019).Article 

    Google Scholar 
    Palagonia, M. S., Brogioli, D. & La Mantia, F. Lithium recovery from diluted brine by means of electrochemical ion exchange in a flow-through-electrodes cell. Desalination 475, 114192 (2020).Article 

    Google Scholar 
    Palagonia, M. S., Brogioli, D. & Mantia, F. L. Effect of current density and mass loading on the performance of a flow-through electrodes cell for lithium recovery. J. Electrochem. Soc. 166, E286–E292 (2019).Article 

    Google Scholar 
    Guo, Z.-Y. et al. Prefractionation of LiCl from concentrated seawater/salt lake brines by electrodialysis with monovalent selective ion exchange membranes. J. Clean. Prod. 193, 338–350 (2018).Article 

    Google Scholar 
    Zhao, L.-M. et al. Separating and recovering lithium from brines using selective-electrodialysis: sensitivity to temperature. Chem. Eng. Res. Des. 140, 116–127 (2018).Article 

    Google Scholar 
    Sharma, P. P. et al. Sulfonated poly (ether ether ketone) composite cation exchange membrane for selective recovery of lithium by electrodialysis. Desalination 496, 114755 (2020).Article 

    Google Scholar 
    Chen, Q.-B. et al. Development of recovering lithium from brines by selective-electrodialysis: effect of coexisting cations on the migration of lithium. J. Memb. Sci. 548, 408–420 (2018).Article 

    Google Scholar 
    Díaz Nieto, C. H. & Flexer, V. Is it possible to recover lithium compounds from complex brines employing electromembrane processes exclusively? Curr. Opin. Electrochem. 35, 101087 (2022).Article 

    Google Scholar 
    Li, X. et al. Highly selective separation of lithium with hierarchical porous lithium-ion sieve microsphere derived from MXene. Desalination 537, 115847 (2022).Article 

    Google Scholar 
    Parker SS, et al. Potential lithium extraction in the United States: environmental, economic, and policy implications (The Nature Conservancy, 2022); https://www.scienceforconservation.org/assets/downloads/Lithium_Report_FINAL.pdf.Arkansas Smackover Project. Standard Lithium https://www.standardlithium.com/projects/arkansas-smackover (2022).Grant, A. Re-injection enhanced production for direct lithium extraction (DLE) projects. https://www.jadecove.com/research/brinereinjection.Horne, R. N. Geothermal reinjection experience in Japan. J. Pet. Technol. 34, 495–503 (1982).Article 

    Google Scholar 
    Boo, C., Billinge, I. H., Chen, X., Shah, K. M. & Yin Yip, N. Zero liquid discharge of ultrahigh-salinity brines with temperature swing solvent extraction. Environ. Sci. Technol. 54, 9124–9131 (2020).Article 

    Google Scholar 
    Deshmukh, A. et al. Thermodynamics of solvent-driven water extraction from hypersaline brines using dimethyl ether. Chem. Eng. J. 434, 134391 (2022).Article 

    Google Scholar 
    Panagopoulos, A. Brine management (saline water & wastewater effluents): sustainable utilization and resource recovery strategy through minimal and zero liquid discharge (MLD & ZLD) desalination systems. Chem. Eng. Process. Process Intensif. 176, 108944 (2022).Article 

    Google Scholar 
    Al-Ghouti, M. A., Al-Kaabi, M. A., Ashfaq, M. Y. & Da’na, D. A. Produced water characteristics, treatment and reuse: a review. J. Water Process Eng. 28, 222–239 (2019).Article 

    Google Scholar 
    Samuel, O. et al. Oilfield-produced water treatment using conventional and membrane-based technologies for beneficial reuse: a critical review. J. Environ. Manag. 308, 114556 (2022).Article 

    Google Scholar 
    Arena, J. T. et al. Management and dewatering of brines extracted from geologic carbon storage sites. Int. J. Greenh. Gas Control 63, 194–214 (2017).Article 

    Google Scholar 
    Ogden, D. D. & Trembly, J. P. Desalination of hypersaline brines via Joule-heating: experimental investigations and comparison of results to existing models. Desalination 424, 149–158 (2017).Article 

    Google Scholar 
    Kaplan, R., Mamrosh, D., Salih, H. H. & Dastgheib, S. A. Assessment of desalination technologies for treatment of a highly saline brine from a potential CO2 storage site. Desalination 404, 87–101 (2017).Article 

    Google Scholar 
    Baspineiro, C. F., Franco, J. & Flexer, V. Potential water recovery during lithium mining from high salinity brines. Sci. Total Environ. 720, 137523 (2020).Article 

    Google Scholar 
    Sustainable production. SQM https://www.sqmlithium.com/en/nosotros/produccion-sustentable/ (2022).Sustainability Report (Orocobre, 2021); https://www.orocobre.com/wp/?mdocs-file=7259.Grant, A. From Catamarca to Qinghai: the Commercial Scale Direct Lithium Extraction Operations. Jadecove https://www.jadecove.com/research/fromcatamarcatoqinghai (2020).Sustainability report (Livent, 2021); https://livent.com/wp-content/uploads/2022/07/Livent_2021SustainabilityReport-English.pdf.Park, S. H. et al. Lithium recovery from artificial brine using energy-efficient membrane distillation and nanofiltration. J. Memb. Sci. 598, 117683 (2020). Pioneering DLE proposal to recover freshwater during brine concentration via nanofiltration followed by membrane distillation.Article 

    Google Scholar 
    Pramanik, B. K., Asif, M. B., Kentish, S., Nghiem, L. D. & Hai, F. I. Lithium enrichment from a simulated salt lake brine using an integrated nanofiltration-membrane distillation process. J. Environ. Chem. Eng. 7, 103395 (2019).Article 

    Google Scholar 
    Ko, C.-C. et al. Performance of ceramic membrane in vacuum membrane distillation and in vacuum membrane crystallization. Desalination 440, 48–58 (2018).Article 

    Google Scholar 
    Baspineiro, C. F., Franco, J. & Flexer, V. Performance of a double-slope solar still for the concentration of lithium rich brines with concomitant fresh water recovery. Sci. Total. Environ. 791, 148192 (2021).Article 

    Google Scholar 
    Ling, Z. et al. Desalination and Li+ enrichment via formation of cyclopentane hydrate. Sep. Purif. Technol. 231, 115921 (2020).Article 

    Google Scholar 
    Niu, J. et al. An electrically switched ion exchange system with self-electrical-energy recuperation for efficient and selective LiCl separation from brine lakes. Sep. Purif. Technol. 274, 118995 (2021).Article 

    Google Scholar 
    Nie, X.-Y., Sun, S.-Y., Song, X. & Yu, J.-G. Further investigation into lithium recovery from salt lake brines with different feed characteristics by electrodialysis. J. Memb. Sci. 530, 185–191 (2017).Article 

    Google Scholar 
    Jiang, C., Wang, Y., Wang, Q., Feng, H. & Xu, T. Production of lithium hydroxide from lake brines through electro-electrodialysis with bipolar membranes (EEDBM). Ind. Eng. Chem. Res. 53, 6103–6112 (2014).Article 

    Google Scholar 
    Díaz Nieto, C. H., Rabaey, K. & Flexer, V. Membrane electrolysis for the removal of Na+ from brines for the subsequent recovery of lithium salts. Sep. Purif. Technol. 252, 117410 (2020).Article 

    Google Scholar 
    Díaz Nieto, C. H. et al. Membrane electrolysis for the removal of Mg2+ and Ca2+ from lithium rich brines. Water Res. 154, 117–124 (2019).Article 

    Google Scholar 
    Ji, P.-Y. et al. Effect of coexisting ions on recovering lithium from high Mg2+/Li+ ratio brines by selective-electrodialysis. Sep. Purif. Technol. 207, 1–11 (2018).Article 

    Google Scholar 
    Lee, D.-H. et al. Selective lithium recovery from aqueous solution using a modified membrane capacitive deionization system. Hydrometallurgy 173, 283–288 (2017).Article 

    Google Scholar 
    Díaz Nieto, C. H., Kortsarz, J. A., Vera, M. L. & Flexer, V. Effect of temperature, current density and mass transport during the electrolytic removal of magnesium ions from lithium rich brines. Desalination 529, 115652 (2022).Article 

    Google Scholar 
    Li, X. et al. Taming wettability of lithium ion sieve via different TiO2 precursors for effective Li recovery from aqueous lithium resources. Chem. Eng. J. 392, 123731 (2020).Article 

    Google Scholar 
    Sun, J. et al. Preparation of high hydrophilic H2TiO3 ion sieve for lithium recovery from liquid lithium resources. Chem. Eng. J. 453, 139485 (2023).Article 

    Google Scholar 
    Qian, F. et al. Trace doping by fluoride and sulfur to enhance adsorption capacity of manganese oxides for lithium recovery. Mater. Des. 194, 108867 (2020).Article 

    Google Scholar 
    Taghvaei, N., Taghvaei, E. & Askari, M. Synthesis of anodized TiO2 nanotube arrays as ion sieve for lithium extraction. ChemistrySelect 5, 10339–10345 (2020).Article 

    Google Scholar 
    Qian, F. et al. Highly lithium adsorption capacities of H1.6Mn1.6O4 ion-sieve by ordered array structure. ChemistrySelect 4, 10157–10163 (2019).Article 

    Google Scholar 
    Sarmiento, N. et al. A solar irradiation GIS as decision support tool for the Province of Salta, Argentina. Renew. Energy 132, 68–80 (2019).Article 

    Google Scholar 
    Dellicompagni, P., Franco, J. & Flexer, V. CO2 emission reduction by integrating concentrating solar power into lithium mining. Energy Fuels 35, 15879–15893 (2021).Article 

    Google Scholar 
    Zavahir, S. et al. A review on lithium recovery using electrochemical capturing systems. Desalination 500, 114883 (2021).Article 

    Google Scholar 
    Joo, H. et al. Pilot-scale demonstration of an electrochemical system for lithium recovery from the desalination concentrate. Environ. Sci. Water Res. Technol. 6, 290–295 (2020). Pilot-scale demonstration of processing 6 tonnes of brine daily using electrochemical ion pumping.Article 

    Google Scholar 
    Song, J. F., Nghiem, L. D., Li, X.-M. & He, T. Lithium extraction from Chinese salt-lake brines: opportunities, challenges, and future outlook. Environ. Sci. Water Res. Technol. 3, 593–597 (2017). Thorough review of pilot-scale projects for lithium mining from brines in China.Article 

    Google Scholar 
    Yu, C. et al. Bio-inspired fabrication of ester-functionalized imprinted composite membrane for rapid and high-efficient recovery of lithium ion from seawater. J. Colloid Interf. Sci. 572, 340–353 (2020).Article 

    Google Scholar 
    Lu, J. et al. Multilayered ion-imprinted membranes with high selectivity towards Li+ based on the synergistic effect of 12-crown-4 and polyether sulfone. Appl. Surf. Sci. 427, 931–941 (2018).Article 

    Google Scholar 
    Ryu, T. et al. Lithium recovery system using electrostatic field assistance. Hydrometallurgy 151, 78–83 (2015).Article 

    Google Scholar 
    Sun, Y., Wang, Y., Liu, Y. & Xiang, X. Highly efficient lithium extraction from brine with a high sodium content by adsorption-coupled electrochemical technology. ACS Sustain. Chem. Eng. 9, 11022–11031 (2021).Article 

    Google Scholar 
    Torres, W. R., Díaz Nieto, C. H., Prévoteau, A., Rabaey, K. & Flexer, V. Lithium carbonate recovery from brines using membrane electrolysis. J. Memb. Sci. 615, 118416 (2020). A DLE methodology with three consecutive electromembrane processes for the sequential recovery of magnesium, calcium and sodium by-products, together with lithium carbonate and concomitant fresh-water production in a circular economy framework.Article 

    Google Scholar 
    Du, X. et al. A novel electroactive λ-MnO2/PPy/PSS core–shell nanorod coated electrode for selective recovery of lithium ions at low concentration. J. Mater. Chem. A 4, 13989–13996 (2016).Article 

    Google Scholar 
    Luo, G. et al. Electrochemical lithium ions pump for lithium recovery from brine by using a surface stability Al2O3–ZrO2 coated LiMn2O4 electrode. J. Energy Chem. 69, 244–252 (2022).Article 

    Google Scholar 
    Oyarce, E., Roa, K., Boulett, A., Salazar-Marconi, P. & Sánchez, J. Removal of lithium ions from aqueous solutions by an ultrafiltration membrane coupled to soluble functional polymer. Sep. Purif. Technol. 288, 120715 (2022).Article 

    Google Scholar 
    Han, H. J., Qu, W., Zhang, Y. L., Lu, H. D. & Zhang, C. L. Enhanced performance of Li+ adsorption for H1.6Mn1.6O4 ion-sieves modified by Co doping and micro array morphology. Ceram. Int. 47, 21777–21784 (2021).Article 

    Google Scholar 
    Zhu, X. et al. Study on adsorption extraction process of lithium ion from West Taijinar brine by shaped titanium-based lithium ion sieves. Sep. Purif. Technol. 274, 119099 (2021).Article 

    Google Scholar 
    Meng, Z. et al. Highly flexible interconnected Li+ion-sieve porous hydrogels with self-regulating nanonetwork structure for marine lithium recovery. Chem. Eng. J. 445, 136780 (2022).Article 

    Google Scholar 
    Xiong, J., Zhao, Z., Liu, D. & He, L. Direct lithium extraction from raw brine by chemical redox method with LiFePO4/FePO4 materials. Sep. Purif. Technol. 290, 120789 (2022).Article 

    Google Scholar 
    Vera, M. L. et al. A strategy to avoid solid formation within the reactor during magnesium and calcium electrolytic removal from lithium-rich brines. J. Solid State Electrochem. https://doi.org/10.1007/s10008-022-05219-6 (2022).Article 

    Google Scholar 
    Lide, D. R. CRC Handbook of Chemistry and Physics (CRC Press. Boca Raton, 2005).Neumann, J. et al. Recycling of lithium-ion batteries — current state of the art, circular economy, and next generation recycling. Adv. Energy Mater. 12, 2102917 (2022).Article 

    Google Scholar 
    Amici, J. et al. A roadmap for transforming research to invent the batteries of the future designed within the European large scale research initiative BATTERY 2030+. Adv. Energy Mater. 12, 2102785 (2022).Article 

    Google Scholar 
    Graedel, T. E. et al. What do we know about metal recycling rates? J. Ind. Ecol. 15, 355–366 (2011).Article 

    Google Scholar 
    Kinnunen, P. H.-M. & Kaksonen, A. H. Towards circular economy in mining: opportunities and bottlenecks for tailings valorization. J. Clean. Prod. 228, 153–160 (2019).Article 

    Google Scholar 
    Falagán, C., Grail, B. M. & Johnson, D. B. New approaches for extracting and recovering metals from mine tailings. Miner. Eng. 106, 71–78 (2017).Article 

    Google Scholar 
    Nwaila, G. T. et al. Valorisation of mine waste — part I: characteristics of, and sampling methodology for, consolidated mineralised tailings by using Witwatersrand gold mines (South Africa) as an example. J. Environ. Manage. 295, 113013 (2021).Article 

    Google Scholar 
    Singh, S., Sukla, L. B. & Goyal, S. K. Mine waste & circular economy. Mater. Today Proc. 30, 332–339 (2020).Article 

    Google Scholar 
    Purnell, P., Velenturf, A. P. M. & Marshall, R. New Governance for Circular Economy: Policy, Regulation and Market Contexts for Resource Recovery from Waste. (eds. Macaskie, L. E. et al.) Ch. 16 (Royal Society of Chemistry, 2019).Velenturf, A. P. M. et al. Circular economy and the matter of integrated resources. Sci. Total Environ. 689, 963–969 (2019).Article 

    Google Scholar 
    Yang, Y. et al. Research advances in magnesium and magnesium alloys worldwide in 2020. J. Magnes. Alloy. 9, 705–747 (2021).Article 

    Google Scholar 
    Løvik, A. N., Hagelüken, C. & Wäger, P. Improving supply security of critical metals: current developments and research in the EU. Sustain. Mater. Technol. 15, 9–18 (2018).
    Google Scholar 
    Potash. Albemarle https://www.albemarle.com/businesses/lithium/products/pot-ash (2023).Production processes. SQM https://www.sqm.com/en/acerca-de-sqm/recursos-naturales/proceso-de-produccion/ (2018).Thiel, G. P. & Lienhard V, J. H. Treating produced water from hydraulic fracturing: composition effects on scale formation and desalination system selection. Desalination 346, 54–69 (2014).Article 

    Google Scholar 
    Shaffer, D. L. et al. Desalination and reuse of high-salinity shale gas produced water: drivers, technologies, and future directions. Environ. Sci. Technol. 47, 9569–9583 (2013).Article 

    Google Scholar 
    Ogunbiyi, O. et al. Sustainable brine management from the perspectives of water, energy and mineral recovery: a comprehensive review. Desalination 513, 115055 (2021).Article 

    Google Scholar 
    Kumar, A. et al. Metals recovery from seawater desalination brines: technologies, opportunities, and challenges. ACS Sustain. Chem. Eng. 9, 7704–7712 (2021).Article 

    Google Scholar 
    Snæbjörnsdóttir, S. Ó. et al. Carbon dioxide storage through mineral carbonation. Nat. Rev. Earth Environ. 1, 90–102 (2020).Article 

    Google Scholar 
    Saccò, M. et al. Salt to conserve: a review on the ecology and preservation of hypersaline ecosystems. Biol. Rev. 96, 2828–2850 (2021).Article 

    Google Scholar 
    Chiappero, M. F., Vaieretti, M. V. & Izquierdo, A. E. A baseline soil survey of two peatlands associated with a lithium-rich salt flat in the argentine puna: physico-chemical characteristics, carbon storage and biota. Mires Peat 27, 16 (2021).
    Google Scholar 
    Batanero, G. L. et al. Flamingos and drought as drivers of nutrients and microbial dynamics in a saline lake. Sci. Rep. 7, 12173 (2017).Article 

    Google Scholar 
    Avila-Arias, H., Nies, L. F., Gray, M. B. & Turco, R. F. Impacts of molybdenum-, nickel-, and lithium-oxide nanomaterials on soil activity and microbial community structure. Sci. Total. Environ. 652, 202–211 (2019).Article 

    Google Scholar 
    Robinson, B. H., Yalamanchali, R., Reiser, R. & Dickinson, N. M. Lithium as an emerging environmental contaminant: mobility in the soil–plant system. Chemosphere 197, 1–6 (2018).Article 

    Google Scholar 
    Bolan, N. et al. From mine to mind and mobiles — lithium contamination and its risk management. Environ. Pollut. 290, 118067 (2021).Article 

    Google Scholar 
    Shokri-Kuehni, S. M. S., Norouzi Rad, M., Webb, C. & Shokri, N. Impact of type of salt and ambient conditions on saline water evaporation from porous media. Adv. Water Resour. 105, 154–161 (2017).Article 

    Google Scholar 
    Obaya, M., López, A. & Pascuini, P. Curb your enthusiasm. Challenges to the development of lithium-based linkages in Argentina. Resour. Policy 70, 101912 (2021).Article 

    Google Scholar 
    Risacher, F., Alonso, H. & Salazar, C. The origin of brines and salts in Chilean salars: a hydrochemical review. Earth-Sci. Rev. 63, 249–293 (2003).Article 

    Google Scholar 
    Corenthal, L. G., Boutt, D. F., Hynek, S. A. & Munk, L. A. Regional groundwater flow and accumulation of a massive evaporite deposit at the margin of the Chilean Altiplano. Geophys. Res. Lett. 43, 8017–8025 (2016).Article 

    Google Scholar 
    Vásquez, C., Ortiz, C., Suárez, F. & Muñoz, J. F. Modeling flow and reactive transport to explain mineral zoning in the Atacama salt flat aquifer, Chile. J. Hydrol. 490, 114–125 (2013).Article 

    Google Scholar  More

  • in

    The determinants of household water consumption: A review and assessment framework for research and practice

    Overview of paper search outcomeA general overview of the 231 general water consumption-related and 48 framework analysis scientific publications reviewed in this study (Fig. 3) shows that the number of papers published per year from the general water consumption-related set has been increasing, particularly since the early 2000s. Peaks of more than 10 papers per year in this category emerge since 2011, with a maximum peak of 34 papers recorded in 2018. This increasing trend in time can be attributed to the increasing development of smart metering studies, which have been increasingly allowing detailed household water demand/consumption and behavioral analysis20,47. As a selected subset of the general water consumption-related papers set, the number of framework analysis papers has also increased in the last decade, compared to the ’80s and ’90s, constituting up to 5 papers per year. The final set of papers includes small-case studies comprising only a few units (11 individual households are considered as a minimum in48), as well as large-scale studies comprising several thousands of households (e.g., more than 8000 individual households are considered in49), or entire communities/towns50.Fig. 3: Temporal development of the literature on the determinants of household water consumption.The yearly count of the 231 general water consumption-related (blue) and 48 framework analysis (orange) scientific publications reviewed in this study is represented for the last forty years.Full size imageFigure 4 shows the locations of the studies from the framework analysis set, with larger blue dots indicating more studies. The geographical distribution of the reviewed studies indicates that the interest in the determinants of water use is worldwide. Prominent interest emerges in particular areas, such as the US west coast, the east coast of Australia, and the Mediterranean area in Europe, perhaps reflecting the combination of areas more prone to drought and/or having the higher economic capacity to undertake water use related research.Fig. 4: Geographical locations of the 48 framework analysis papers.The location of the 48 framework analysis papers reviewed in this study is represented with blue markers. Marker size is proportional to the amount of studies in a specific location.Full size imageDeterminant representation by classFrom the analysis of the 48 framework analysis papers, we identified a range of heterogeneous determinants and quantified different combinations of determinant classes, namely observable, latent, and external (see Determinant classification). Figure 5 shows an overview of the representation for the different classes of determinants over the 48 analyzed papers. Observable determinants were the most popular (47 total studies, i.e., 98% representation), with latent and external having lower representation of 52% and 56%, respectively. The values represented in the figure confirm our hypotheses that observable determinants are more common in literature than latent determinants, due to their availability in public databases, either at the household level or at coarser spatial sub-urban scales (e.g., census data collected at the block group-level, such as those used in49). The slightly higher representation of external determinants, compared to latent determinants, is also as expected due to the widespread availability of weather records (e.g., temperature, rainfall) from national or international environmental agencies. While there is no full consensus in the literature on the effect of weather or price variables on water consumption51,52,53, the high degree of representation of external determinants demonstrates that they are considered in more than half of the studies.Fig. 5: Venn diagram of household water consumption determinants representation.The representation of different classes of determinants (observable, latent, and external) in the 48 reviewed framework analysis papers is represented with coloured circles. Intersections are also visualized. The size of each circle and the numerical labels indicate the number of studies in which each combination of determinant classes appeared.Full size imageIt is worth observing that multiple classes of determinants are simultaneously analyzed in most of the reviewed studies, with fewer than 20% analyzing observable variables alone. Further, almost every time external or latent variables are considered, they appear in combination with observable variables. Only one study specifically focused on analyzing the motivations for using and conserving water based on only latent determinants42, while no studies exclusively considered external variables. In contrast, nearly 30% of the studies included both observable and external variables, approximately 23% of the studies considered latent and external variables simultaneously, and 27% of the studies included all three types of determinant classes.The high representation of observable determinants (Fig. 5) suggests that observable variables are widespread in the literature on modelling and forecasting of household water consumption. The prevalence of this class of determinants seems also to confirm the findings from previous studies, which demonstrated that meteorological variables have a greater influence on medium-term prediction and urban/suburban scales, but socio-demographics become more relevant when household-scale and short-term water demand models are developed54,55.Individual determinant representation, impact, and effortTo facilitate interpretation of the numerical values we obtained for the three determinant assessment criteria (i.e., representation, impact, and effort) we defined some regions of interest for each criterion based on thresholds (see the regions labelled as low/high/very high in Fig. 6). We selected the threshold values used to delimit the above regions of interest based on visual inspection of the empirical distribution of the representation, impact, and effort values. This simplification is carried out to facilitate the inference of general qualitative conclusions, while accounting for the low number of papers and, at the same time, high number of determinants. As a result, representation values above/below 30% are considered high/low. Impact values below 75% are considered low, values between 75% and 90% are considered high, and values above 90% are considered very high. Effort rate values above/below 8 are considered high/low.Fig. 6: Individual analysis of representation, impact, and effort.The three criteria to perform determinant analysis, i.e., representation (top), impact (middle), and effort rate (bottom) are associated with individual determinants. Observable determinant class is shown in green, latent class in blue, and external in orange. Shaded background indicates different levels of intensity for each analysis criterion. See Tables 1–3 for determinant acronyms definition (the determinants included in the categories marked as “Other” in the tables are not represented for better clarity).Full size imageFrom the resulting data visualized in Fig. 6, we can infer the following insights about determinant representation, impact, and effort. First, the determinants with the highest representation (top plot in Fig. 6) were household income ( > 70%), family size ( > 60%), and age ( > 45%). As already suggested by the outcomes of class representation (Fig. 5), all the above determinants with high representation are observable. One exception is the awareness determinant, which is the only non-observable determinant we found with high representation. The majority of the other determinants had a representation rate of 10% to 30%.Second, the number of determinants with a high or very high impact (middle plot in Fig. 6) is much larger than the number of determinants with high representation. It must be noted that a high impact does not necessarily mean that a determinant was found to have a high influence on water consumption, but rather that it was found to have some influence on water consumption in many publications. Interestingly, some determinants from all classes achieve high or very high levels of impact. Observable determinants with very high impact include socio-demographic information (number of occupants), house characteristics (house age, value), and outdoor characteristics (garden size, and presence of rainwater tanks). While these latter attributes related to gardens ranked among those with the highest impact, garden composition was found to have one of the lowest impact rates across the analyzed studies. Also, the observable determinants related to the education level of occupants was found to have low impact. A latent variable that emerges as very important (GARD_C) is also related to garden characteristics, but, rather than representing any physical variable, it accounts for the psychological value given by occupants’ attitudes and habits towards gardening. Finally, all external variables were found to have high or very high impact, with rainfall and water price emerging as the two with impact above 90%.The bottom plot of Fig. 6 shows that there was a wide variability in the effort rate for each individual determinant. Data on most of the observable determinants can be generally gathered with low effort, but some (e.g., appliance inventory and irrigation system) require house visits, and thus require high effort. In turn, all latent variables display an effort rate higher than 6, and three out of four are classified as high-effort. Conversely, data on all external determinants can be retrieved with low effort, as they are usually available from national agencies (weather data) and water utilities (water price). Obtaining information on higher effort determinants likely requires getting in contact with individual householders, via phone/online surveys, or house visits.Overall, the results reported in Fig. 6 suggest that there are trade-offs between representation, impact, and effort. In the next section, we perform a joint analysis of the three criteria and their trade-off to infer the implications of the outcomes of this study for researchers and practitioners.Trade-off analysis and implications for researchers and practitionersFigure 7 shows the interaction between the representation, impact, and effort criteria. The distribution of blue and orange points in the figure demonstrates that there are different trade-offs among the three criteria. Each trade-off can have a different set of implications to derive recommendations for researchers and also practitioners. We identified the three groups of determinants marked with (A), (B), and (C) to illustrate the different needs of research and practice. Group A is characterized by high impact, high representation, and low effort. Determinants in this group include household family size, occupants’ age, and occupants’ income. This group of well-studied determinants with proven impact might be particularly interesting for practitioners aiming to gather knowledge on household water consumption with budget constraints. Group (B), which includes, among others, information on the household irrigation system, appliance efficiency, and occupant gender, is characterized by medium-to-high impact, but low representation, and a range of low to high effort. While this group might not be very appealing for practitioners due to low representation, researchers might be interested in focusing on these determinants to increase their representation and, thus, validate or contrast the limited findings on these determinants that appear in the literature. Finally, Group C refers to determinants with low representation and, compared to those in groups A and B, lower impact. As they also might require high gathering efforts, these determinants should be treated with caution until more research is performed to prove their potential impact on a larger sample of studies.Fig. 7: Trade-off analysis of determinant representation, impact, and effort.Impact (x-axis) vs Representation (y-axis) vs Effort (color) of each determinant. Each point refers to a specific determinant. See Tables 1–3 for determinant acronyms definition. The determinants classified as “High effort” are those with an effort value larger than 8.0, vice-versa for the “Low effort” determinants. Determinants are organized in three groups: Group A – high impact, high representation, and low effort; Group B – medium-to-high impact, low representation, and mixed low and high effort; Group C – low representation, low impact, mixed effort.Full size imageAccounting for similar trade-offs across the entire sample of determinants that we have identified from the review of the literature enables determinant-specific recommendations to be derived for practitioners and researchers. In the last step of this review and determinant classification effort we thus develop a trade-off analysis framework that considers different combinations of representation, effort, and impact to formulate such recommendations. In keeping with the goal of this study, our trade-off analysis aims at identifying groups of determinants that have proven cost-effective impact via extensive research and, thus, can be recommended for use in practice, compared with groups of determinants that require more research to address open questions related to representation, impact, and effort. The proposed trade-off analysis framework includes four main recommendation categories:UIn this category, we include determinants characterized by high representation, high/very high impact, and low effort. We consider these determinants as determinants that practitioners can “definitely use” (U), as they have been extensively researched and have been shown to have an impact in most cases, while being affordable. For the same reasons, higher levels of research priority should be devoted to less explored determinants, while these can serve as references. The determinants included in box (A) in Fig. 7 belong to this group.IR-UCIn this category, we classify those determinants characterized by low representation, high/very high impact, and low effort. Given their promising, but not extensively proven, impact, and overall affordability, further research on these determinants should be prioritized to increase their representation (IR). We consider these determinants as determinants that practitioners can “use with caution” (UC), as they have not been extensively researched, but at the same time might have high impact at low-cost. The determinants included in box (B) in Fig. 7 and classified as low effort (blue color) belong to this group.LE/IR-UCIn this category, we include determinants characterized by generally low representation, high/very high impact, and high effort. Similarly to the previous category, we believe that practitioners can use these determinants “with caution” (UC), as they have not been extensively researched and require high effort for data collection, but at the same time might have high/very high impact. Given their promising, but not extensively proven, impact, and high cost, further research on these determinants should be prioritized, primarily to lower the effort (LE) needed to collect them and, thus, facilitate their consideration in more studies (increased representation – IR). The determinants included in box (B) in Fig. 7 and classified as high effort (orange color) belong to this group.IR/LE/AI-NPIn this category, we include determinants characterized by low representation, low impact, and mainly high effort. Given the limited knowledge on these determinants, we suggest that these determinants are “not prioritized” (NP) for use by practitioners unless further research demonstrates that the effort required to collect these determinants is worth the benefit of considering them. Further research should then aim at increasing their representation (IR), lowering the effort needed to obtain data on these determinants (LE), and further assessing their impact (AI) to acquire better knowledge on their actual value. The determinants included in box (C) in Fig. 7 belong to this group.Summary information on the above categories is reported in Table 5. Based on the proposed trade-off analysis framework and the threshold values defined in Fig. 6, we associated each of the different determinants identified in the framework analysis papers with a level of recommendation (see Fig. 8). Some relevant insights for researchers and practitioners emerge. First, only observable determinants are classified as “U”. At present, there are some socio-demographic determinants (i.e., number of occupants, income level, and occupant age) that can be reliably used by practitioners in most cases to model household water consumption and can be easily and affordably retrieved.Table 5 Framework for trade-off analysis, based on representation, impact, and effort.Full size tableFig. 8: Household water consumption determinant classification and associated recommendations for practitioners and researchers based on individual determinants.Each household water consumption determinant identified in the framework analysis papers is associated with a level of recommendation. Determinants are classified according to the three defined classes (columns), i.e., observable, latent, and external. Four levels of recommendation (rows) are formulated for practitioners and researchers. They are sorted in decreasing order of representation and proven impact in research, as well as confidence for use in practical applications. Confidence for use in practical applications decreases going from green (“U” level of recommendation) to orange (“IR/LE/AI-NP” level of recommendation).Full size imageSecond, all external variables (i.e., average rainfall, temperature, and water price) are classified as IR-UC. Consequently, they have a proven impact, but have been used sporadically in connection with household water consumption (while they have been used more often at larger, urban scales), thus results might be case-specific and further research is needed to assess their impact on a larger number of studies.Third, a mix of observable and latent external variables deserves further research to lower effort (e.g., by improving technology/data gathering practices or identifying lower-effort proxies for the same type of information) and increase representation. These variables are either observable determinants, the collection of which requires significant effort and house visits/calls to occupants (e.g., to build an inventory of appliance efficiency or storing information on irrigation systems), or latent variables the impact of which is still not proven due to low representation. The increasing availability of high-resolution metering and behavioral studies fostered by smart metering development is likely to contribute more knowledge on these determinants and more complete guidelines for use by practitioners in the coming years7,17,20.Fourth, we would like to stress that the recommendation “Do not favor adoption until further research” for the determinants classified as IR/LE/AI-NP does not mean that they should not be considered in future applications or no research should be done on them. Conversely, we recognize that many existing studies are based on limited data or data with coarser spatio-temporal resolutions, thus conclusive statements on the impact of such determinants would require further validation. Since large uncertainty about their impact remains, more studies are actually needed to increase the representation of these determinants and increase the statistical significance and generality of their impact assessment. Joint research that also includes other determinants with higher levels of representation could be beneficial to discover more information on the determinants in this group and better understand whether practitioners should eventually include one/more of these determinants in their analysis. Further research could be also developed to assess the degree to which these determinants are correlated with others, and hence redundant, and to which extent these and other determinants can relate to particular characteristics of water consumption (e.g., demand peaks, end use components).Finally, some of the determinants that we recommend to use with caution (UC) in practice, and that should be prioritized for research, might become determinants to definitely use (U) in the future. Two limitations currently prevent us to recommend “definitely use” for these determinants, i.e., generally low representation and high effort for data collection. Low representation indicates that the determinant has not been well-studied in the literature. Hence it might not be generalizable to a wide range of locations. To address the disadvantages of low representation, the following is recommended for practitioners:

    Check the literature and if there are studies with similar context (location/climate/application) to the practitioners’ required application, and the impact of the determinant is high, then the determinant could be considered for use.

    Continue to monitor the literature, to see if new studies appear using that determinant.

    The other limitation, high effort, means that in the reviewed past studies it has been costly for practitioners to collect some of the required determinants. With the advent and widespread use of new technologies, the effort required to collect some of the required high-effort determinants may be substantially reduced. Lowering the effort related to some high-impact, yet also high-effort, determinants (see, e.g., those indicated with orange color in Group B in Fig. 7) would have a two-fold benefit. The direct reduction of the costs required to collect information on those determinants will also enable wider consideration of these determinants in a larger number of studies, thus increasing their representation. As technology enhances the capture of such determinants, there is an opportunity to revisit past studies/datasets and increase the representation of these determinants, which might then transition to determinant group A in in Fig. 7. To address the current limitations and disadvantages of high effort, the following is recommended for practitioners:

    Monitor the use of emerging technologies that provide an opportunity to lower the cost required to collect the determinant. For example, there is an opportunity for analysis of high resolution satellite maps/photos to provide automated estimates of observable determinants such as garden size (GSZE) over large number of households, which would lower the cost substantially56. Similarly, latent determinants such as water consumption awareness (AWARE_C) could be based on the uptake of user-friendly smart metering and phone apps on water consumption if they were widely available17,57.

    Evaluate overall costs vs benefits based on preliminary experiments on small sample data (to evaluate benefits while avoiding high costs), and consider the use of lower cost proxy data for the “high effort” determinant.

    These recommendations provide some guidance for practitioners to handle determinants classified as “use with caution”.Limitations and future researchThis work provides evidence and a quantitative framework for the analysis of household water consumption determinants, yet several limitations and questions remain for further research. First, alternative formulations of determinant representations, impact, and effort could lead to different results. This also stands for the subjective thresholds we adopted to distinguish between high and low representation, impact, and effort. Such thresholds and criteria formulation could be changed based on needs and subjective judgement.Second, in this review we focused on the analysis of individual determinants of household water consumption. However, some determinants could be correlated, present redundant information, or be accounted for in alternative ways to build models for forecasting water demand (e.g., rainfall amount versus rainfall occurrence53). Input feature engineering, variable redundancy, and data accuracy can substantially affect the performance of water demand models. Future studies focused on comparative analysis of alternative determinant formulations and inter-links/dependencies among different determinants can help define non-redundant determinant sets to train models of water demand and recommendations for variable pre-processing.Third, the findings of this study are consistent with previous review papers that identified both observable and latent variables as the most important with respect to domestic water consumption58. Yet, other meta-analyses and review studies found partly contradictory results. Differently from our study26, found that the most important determinants of water use behaviour are related to individual opportunities and motivations, gender, income, and education level. In turn59, formulated a model that accounted for a wide range of variables including demographics, dwelling characteristics, household composition, conservation intention, trust, perceptions, habits, and perceived behavioral control. It must be noted that the above studies do not consider household water consumption per se as we do here, but relate potential determinants of water consumption also to individual consumption or behavior changes (i.e., changes in water consumption over time). Future, potentially contrasting, studies could then expand the scope of this work and relax the exclusion criteria we adopted here to achieve more inclusive comparative analyses that investigate the effect of different determinants in relation to quantified intervals of total household water consumption, and other heterogeneous aspects of domestic water demand, including statistics on end use components (e.g., flow rate, duration, or frequency of individual appliances)60 and temporal changes of water consumption levels due to external stressors such as droughts, or demand management interventions39,49, for example.Fourth, the set of framework analysis papers includes case studies primarily located in the United States, Australia, and Europe (see Fig. 4). Geographical coverage is thus skewed. There is a need for more studies from other geographical regions (including countries with low-income economies) in order to obtain a more balanced picture and consolidate/expand the results obtained so far.Finally, recent works have highlighted that urban and household water demands have been modelled at different spatial and temporal resolutions47. The choice of the temporal and spatial resolution of interest is determined both by data availability and the specific modelling and management purpose. Multi-scale studies combining different levels of spatial and temporal aggregation of water demands and potential determinants would further advance our analysis and contextualize specific recommendations for data collection and processing at the different spatial and temporal scales of interest. More

  • in

    A three-dimensional antifungal wooden cone evaporator for highly efficient solar steam generation

    Goh, P. S., Matsuura, T., Ismail, A. F. & Ng, B. C. The water-energy nexus: solutions towards energy-efficient desalination. Energy Technol. 5, 1136–1155 (2017).Article 

    Google Scholar 
    Jia, C., Yan, P., Liu, P. & Li, Z. Energy industrial water withdrawal under different energy development scenarios: a multi-regional approach and a case study of China. Renew. Sust. Energ. Rev. 135, 110224 (2021).Article 

    Google Scholar 
    Damkjaer, S. & Taylor, R. The measurement of water scarcity: defining a meaningful indicator. Ambio 46, 513–531 (2017).Article 

    Google Scholar 
    Venugopal, K. & Dharmalingam, S. Utilization of bipolar membrane electrodialysis for salt water treatment. Water Environ. Res. 85, 663–670 (2013).Article 
    CAS 

    Google Scholar 
    Camacho, L. et al. Advances in membrane distillation for water desalination and purification applications. Water 5, 94–196 (2013).Article 

    Google Scholar 
    Cingolani, D., Eusebi, A. L. & Battistoni, P. Osmosis process for leachate treatment in industrial platform: economic and performances evaluations to zero liquid discharge. J. Environ. Manag. 203, 782–790 (2017).Article 
    CAS 

    Google Scholar 
    Shahzad, M. W., Burhan, M., Ang, L. & Ng, K. C. in Emerging Technologies for Sustainable Desalination Handbook (ed. Gude, V. G.) Ch. 1 (Elsevier Science, 2018)Feria-Díaz, J. J., López-Méndez, M. C., Rodríguez-Miranda, J. P., Sandoval-Herazo, L. C. & Correa-Mahecha, F. Commercial thermal technologies for desalination of water from renewable energies: a state of the art review. Processes 9, 262 (2021).Article 

    Google Scholar 
    Ahmad, N. A., Goh, P. S., Yogarathinam, L. T., Zulhairun, A. K. & Ismail, A. F. Current advances in membrane technologies for produced water desalination. Desalination 493, 114643 (2020).Article 
    CAS 

    Google Scholar 
    Errico, M. et al. Membrane assisted reactive distillation for bioethanol purification. Chem. Eng. Process 157, 108110 (2020).Article 
    CAS 

    Google Scholar 
    Saren, S., Mitra, S., Miyazaki, T., Ng, K. C. & Thu, K. A novel hybrid adsorption heat transformer – multi-effect distillation (AHT-MED) system for improved performance and waste heat upgrade. Appl. Energy 305, 117744 (2022).Article 

    Google Scholar 
    Son, H. S., Shahzad, M. W., Ghaffour, N. & Ng, K. C. Pilot studies on synergetic impacts of energy utilization in hybrid desalination system: multi-effect distillation and adsorption cycle (MED-AD). Desalination 477, 114266 (2020).Shahzad, M. W., Burhan, M. & Ng, K. C. A standard primary energy approach for comparing desalination processes. NPJ Clean. Water 2, 1–7 (2019).Article 
    CAS 

    Google Scholar 
    Zheng, X., Chen, D., Wang, Q. & Zhang, Z. Seawater desalination in China: retrospect and prospect. Chem. Eng. J. 242, 404–413 (2014).Article 
    CAS 

    Google Scholar 
    Werber, J. R., Osuji, C. O. & Elimelech, M. Materials for next-generation desalination and water purification membranes. Nat. Rev. Mater. 1, 1–15 (2016).Article 

    Google Scholar 
    Mahmoud, K. A., Mansoor, B., Mansour, A. & Khraisheh, M. Functional graphene nanosheets: the next generation membranes for water desalination. Desalination 356, 208–225 (2015).Article 
    CAS 

    Google Scholar 
    Homaeigohar, S. & Elbahri, M. Graphene membranes for water desalination. NPG Asia Mater. 9, e427–e427 (2017).Article 
    CAS 

    Google Scholar 
    Li, X. et al. Enhancement of interfacial solar vapor generation by environmental energy. Joule 2, 1331–1338 (2018).Article 
    CAS 

    Google Scholar 
    Brongersma, M. L., Halas, N. J. & Nordlander, P. Plasmon-induced hot carrier science and technology. Nat. Nanotechnol. 10, 25–34 (2015).Article 
    CAS 

    Google Scholar 
    Ni, G. et al. Volumetric solar heating of nanofluids for direct vapor generation. Nano Energy 17, 290–301 (2015).Article 
    CAS 

    Google Scholar 
    Wang, X., Ou, G., Wang, N. & Wu, H. Graphene-based recyclable photo-absorbers for high-efficiency seawater desalination. ACS Appl. Mater. Interfaces 8, 9194–9199 (2016).Article 
    CAS 

    Google Scholar 
    Ghasemi, H. et al. Solar steam generation by heat localization. Nat. Commun. 5, 4449 (2014).Article 
    CAS 

    Google Scholar 
    Han, X. et al. Intensifying heat using MOF‐isolated graphene for solar‐driven seawater desalination at 98% solar‐to‐thermal efficiency. Adv. Funct. Mater. 31, 2008904 (2021).Article 
    CAS 

    Google Scholar 
    Luo, X. et al. The energy efficiency of interfacial solar desalination. Appl. Energy 302, 117581 (2021).Article 

    Google Scholar 
    Mahian, O., Kianifar, A., Kalogirou, S. A., Pop, I. & Wongwises, S. A review of the applications of nanofluids in solar energy. Int. J. Heat. Mass Transf. 57, 582–594 (2013).Article 
    CAS 

    Google Scholar 
    Ni, G. et al. Steam generation under one sun enabled by a floating structure with thermal concentration. Nat. Energy 1, 1–7 (2016).Article 

    Google Scholar 
    Panchal, H., Patel, P., Patel, N. & Thakkar, H. Performance analysis of solar still with different energy-absorbing materials. Int. J. Ambient. Energy 38, 224–228 (2015).Article 

    Google Scholar 
    Li, S.-F., Liu, Z.-H., Shao, Z.-X., Xiao, H.-S. & Xia, N. Performance study on a passive solar seawater desalination system using multi-effect heat recovery. Appl. Energy 213, 343–352 (2018).Article 

    Google Scholar 
    Yin, Z. et al. Extremely black vertically aligned carbon nanotube arrays for solar steam generation. ACS Appl. Mater. Interfaces 9, 28596–28603 (2017).Article 
    CAS 

    Google Scholar 
    Wang, Z. et al. A wood–polypyrrole composite as a photothermal conversion device for solar evaporation enhancement. J. Mater. Chem. A 7, 20706–20712 (2019).Article 
    CAS 

    Google Scholar 
    Ghim, D., Jiang, Q., Cao, S., Singamaneni, S. & Jun, Y.-S. Mechanically interlocked 1T/2H phases of MoS2 nanosheets for solar thermal water purification. Nano Energy 53, 949–957 (2018).Article 
    CAS 

    Google Scholar 
    Ito, Y. et al. Multifunctional porous graphene for high-efficiency steam generation by heat localization. Adv. Mater. 27, 4302–4307 (2015).Article 
    CAS 

    Google Scholar 
    Li, T., Fang, Q., Xi, X., Chen, Y. & Liu, F. Ultra-robust carbon fibers for multi-media purification via solar-evaporation. J. Mater. Chem. A 7, 586–593 (2019).Article 
    CAS 

    Google Scholar 
    Liu, H. et al. Narrow bandgap semiconductor decorated wood membrane for high-efficiency solar-assisted water purification. J. Mater. Chem. A 6, 18839–18846 (2018).Article 
    CAS 

    Google Scholar 
    Wang, G. et al. Reusable reduced graphene oxide based double-layer system modified by polyethylenimine for solar steam generation. Carbon 114, 117–124 (2017).Article 
    CAS 

    Google Scholar 
    Yang, Y. et al. Two-dimensional flexible bilayer janus membrane for advanced photothermal water desalination. ACS Energy Lett. 3, 1165–1171 (2018).Article 
    CAS 

    Google Scholar 
    Wang, Y., Wu, X., Yang, X., Owens, G. & Xu, H. Reversing heat conduction loss: extracting energy from bulk water to enhance solar steam generation. Nano Energy 78, 105269 (2020).Article 
    CAS 

    Google Scholar 
    Zhu, M. et al. Plasmonic wood for high-efficiency solar steam generation. Adv. Energy Mater. 8, 1701028 (2018).Article 

    Google Scholar 
    Gong, F. et al. Scalable, eco-friendly and ultrafast solar steam generators based on one-step melamine-derived carbon sponges toward water purification. Nano Energy 58, 322–330 (2019).Article 
    CAS 

    Google Scholar 
    Kong, Y. et al. Self-floating maize straw/graphene aerogel synthesis based on microbubble and ice crystal templates for efficient solar-driven interfacial water evaporation. J. Mater. Chem. A 8, 24734–24742 (2020).Article 
    CAS 

    Google Scholar 
    Wang, M., Wang, P., Zhang, J., Li, C. & Jin, Y. A ternary Pt/Au/TiO2 -decorated plasmonic wood carbon for high-efficiency interfacial solar steam generation and photodegradation of tetracycline. ChemSusChem 12, 467–472 (2019).Article 
    CAS 

    Google Scholar 
    Li, T. et al. Scalable and highly efficient mesoporous wood-based solar steam generation device: localized heat, rapid water transport. Adv. Funct. Mater. 28, 1707134 (2018).Article 

    Google Scholar 
    Liu, K. K. et al. Wood-graphene oxide composite for highly efficient solar steam generation and desalination. ACS Appl. Mater. Interfaces 9, 7675–7681 (2017).Article 
    CAS 

    Google Scholar 
    Zou, Y. et al. Boosting solar steam generation by photothermal enhanced polydopamine/wood composites. Polymer 217, 123464 (2021).Article 
    CAS 

    Google Scholar 
    Li, Y. et al. 3D-printed, all-in-one evaporator for high-efficiency solar steam generation under 1 Sun illumination. Adv. Mater. 29, 1700981 (2017).Article 

    Google Scholar 
    Hong, S. et al. Nature-inspired, 3D origami solar steam generator toward near full utilization of solar energy. ACS Appl. Mater. Interfaces 10, 28517–28524 (2018).Article 
    CAS 

    Google Scholar 
    Zhang, L., Li, R., Tang, B. & Wang, P. Solar-thermal conversion and thermal energy storage of graphene foam-based composites. Nanoscale 8, 14600–14607 (2016).Article 
    CAS 

    Google Scholar 
    Tu, C. et al. A 3D-structured sustainable solar-driven steam generator using super-black nylon flocking materials. Small 15, e1902070 (2019).Article 

    Google Scholar 
    Pham, T. T. et al. Durable, scalable and affordable iron (III) based coconut husk photothermal material for highly efficient solar steam generation. Desalination 518, 115280 (2021).Article 
    CAS 

    Google Scholar 
    Shao, B. et al. A general method for selectively coating photothermal materials on 3D porous substrate surfaces towards cost-effective and highly efficient solar steam generation. J. Mater. Chem. A 8, 24703–24709 (2020).Article 
    CAS 

    Google Scholar 
    Yuan, B. et al. A low‐cost 3D spherical evaporator with unique surface topology and inner structure for solar water evaporation‐assisted dye wastewater treatment. Adv. Sustain. Syst. 5, 2000245 (2020).Article 

    Google Scholar 
    Xu, Y. et al. Origami system for efficient solar driven distillation in emergency water supply. Chem. Eng. J. 356, 869–876 (2019).Article 
    CAS 

    Google Scholar 
    Kim, K., Yu, S., Kang, S.-Y., Ryu, S.-T. & Jang, J.-H. Three-dimensional solar steam generation device with additional non-photothermal evaporation. Desalination 469, 114091 (2019).Article 
    CAS 

    Google Scholar 
    Shi, Y. et al. A 3D photothermal structure toward improved energy efficiency in solar steam generation. Joule 2, 1171–1186 (2018).Article 
    CAS 

    Google Scholar 
    Sui, Y., Hao, D., Guo, Y., Cai, Z. & Xu, B. A flowerlike sponge coated with carbon black nanoparticles for enhanced solar vapor generation. J. Mater. Sci. 55, 298–308 (2019).Article 

    Google Scholar 
    Cao, N. et al. A self-regenerating air-laid paper wrapped ASA 3D cone-shaped Janus evaporator for efficient and stable solar desalination. Chem. Eng. J. 397, 125522 (2020).Article 
    CAS 

    Google Scholar 
    Wang, Y. et al. Improved light-harvesting and thermal management for efficient solar-driven water evaporation using 3D photothermal cones. J. Mater. Chem. A 6, 9874–9881 (2018).Article 
    CAS 

    Google Scholar 
    Liu, H. et al. Conformal microfluidic-blow-spun 3D photothermal catalytic spherical evaporator for omnidirectional enhanced solar steam generation and CO2 reduction. Adv. Sci. 8, e2101232 (2021).Article 

    Google Scholar 
    Gong, Y. et al. Towards suppressing dielectric loss of GO/PVDF nanocomposites with TA-Fe coordination complexes as an interface layer. J. Mater. Sci. Technol. 34, 2415–2423 (2018).Article 
    CAS 

    Google Scholar 
    Mehrkhah, R., Goharshadi, E. K. & Mohammadi, M. Highly efficient solar desalination and wastewater treatment by economical wood-based double-layer photoabsorbers. J. Ind. Eng. Chem. 101, 334–347 (2021).Article 
    CAS 

    Google Scholar 
    Kuang, Y. et al. A high-performance self-regenerating solar evaporator for continuous water desalination. Adv. Mater. 31, e1900498 (2019).Article 

    Google Scholar 
    Guan, H., Cheng, Z. & Wang, X. Highly compressible wood sponges with a spring-like lamellar structure as effective and reusable oil absorbents. ACS Nano 12, 10365–10373 (2018).Article 
    CAS 

    Google Scholar 
    Ghafurian, M. M. et al. Enhanced solar desalination by delignified wood coated with bimetallic Fe/Pd nanoparticles. Desalination 493, 114657 (2020).Article 
    CAS 

    Google Scholar 
    Sahiner, N., Butun Sengel, S. & Yildiz, M. A facile preparation of donut-like supramolecular tannic acid-Fe(III) composite as biomaterials with magnetic, conductive, and antioxidant properties. J. Coord. Chem. 70, 3619–3632 (2017).Article 
    CAS 

    Google Scholar 
    Huang, Y., Lin, Q., Yu, Y. & Yu, W. Functionalization of wood fibers based on immobilization of tannic acid and in situ complexation of Fe (II) ions. Appl. Surf. Sci. 510, 145436 (2020).Article 
    CAS 

    Google Scholar 
    Song, L., Zhang, X.-F., Wang, Z., Zheng, T. & Yao, J. Fe3O4/polyvinyl alcohol decorated delignified wood evaporator for continuous solar steam generation. Desalination 507, 115024 (2021).Article 
    CAS 

    Google Scholar 
    Li, X. et al. Graphene oxide-based efficient and scalable solar desalination under one sun with a confined 2D water path. Proc. Natl Acad. Sci. 113, 13953–13958 (2016).Article 
    CAS 

    Google Scholar 
    Li, X. Q. et al. Three-dimensional artificial transpiration for efficient solar waste-water treatment. Natl Sci. Rev. 5, 70–77 (2018).Article 
    CAS 

    Google Scholar 
    Ma, M., Dong, S., Hussain, M. & Zhou, W. Effects of addition of condensed tannin on the structure and properties of silk fibroin film. Polym. Int. 66, 151–159 (2017).Article 
    CAS 

    Google Scholar 
    Jiang, P. et al. Synthesis of flame-retardant, bactericidal, and color-adjusting wood fibers with metal phenolic networks. Ind. Crops Prod. 170, 113796 (2021).Article 
    CAS 

    Google Scholar 
    Song, J. et al. Processing bulk natural wood into a high-performance structural material. Nature 554, 224–228 (2018).Article 
    CAS 

    Google Scholar  More

  • in

    Hauling icebergs to Africa: could a bizarre plan to get drinking water actually work?

    Chasing Icebergs: How Frozen Freshwater Can Save the Planet Matthew H. Birkhold Viking (2023)The flavour of chilled Svalbarði-brand water, melted from an iceberg just 1,000 kilometres from the North Pole, is described by the company as “like catching snowflakes on the tongue”. Bottled in Longyearbyen, a tiny metropolis in Norway’s Svalbard archipelago, Svalbarði water is airlifted to luxury locales in London, Sydney, Florida and Macau. “Taste the Arctic to Save the Arctic,” its website croons, promoting the supposed carbon neutrality of the water, which sells online for €99.95 (US$107) for a 750-millilitre bottle.That price is far out of reach for most of the world’s people, including the one in four who lack safe drinking water. In Chasing Icebergs, Matthew Birkhold, a scholar of law, culture and the humanities, considers whether it’s possible to slake the world’s thirst with the two-thirds of global fresh water that is locked away in ice caps and glaciers — “stuck at the poles in gigantic fortresses of ice”, in his words.Some are already at it — and not just epicureans quaffing Svalbarði. In Newfoundland, Canada, Birkhold interviews “iceberg cowboy” Ed Kean, who wrangles bergs from the frigid sea, selling the water to cosmetics companies and breweries. In Qaanaaq, Greenland’s northernmost town, Birkhold notes, the public water supply includes filtered and treated iceberg melt.
    Towing icebergs to arid regions to reduce water scarcity
    There’s more where that came from. By one estimate, some 2,300 cubic kilometres of ice breaks off from Antarctica every year. More than 100,000 Arctic and Antarctic icebergs melt into the ocean annually, according to a 2022 United Nations report. A relatively small, 113-million-tonne iceberg, says Birkhold, could be towed from Antarctica to Cape Town, South Africa, to supply 20% of the city’s water needs for a year. What’s not to like?Quite a lot, perhaps. “To write this book, I have talked to dozens of scientists,” Birkhold writes. “They are uniformly dubious.” Palaeoclimatologist Ellen Mosley-Thompson has led nine expeditions to Antarctica and six to Greenland to extract ice cores. “Make sure you write that I am skeptical of iceberg towing,” she instructs him.Less sceptical is master mariner Nick Sloane, the brains behind the Cape Town plan. Using satellite data to find the best berg, his “team of glaciologists, engineers and oceanographers” plans to catch it in a giant net and pull it by tugboat into the mighty Antarctic Circumpolar Current, and thence into the north-flowing Benguela Current towards South Africa.

    Icebergs are sometimes towed to prevent damage to oil platforms.Credit: Greg Locke/Reuters

    Sloane has estimated the cost at a cool $100 million, plus an extra $50 million or so to melt the ice and funnel the fresh water to land, if the iceberg hasn’t melted into the sea or fallen apart en route. In various interviews, Cape Town officials have expressed themselves as less than enthused.Sloane’s plans have yet to materialize. Meanwhile, POLEWATER, a Berlin-based company, has been working for almost a decade on a similar plan — to haul frozen fresh water to the western coast of Africa and the Caribbean, where the company intends to give it away to those in dire need. It, too, will use satellites to locate suitable bergs, but once it has relocated them, it plans to pump meltwater from pools on top into easily transportable gigantic bags.Then there’s the UAE Iceberg Project, the dream of Emirati inventor Abdulla Alshehhi to import an Antarctic iceberg to the Fujairah coast of the United Arab Emirates. An animated promotion features penguins and polar bears — species hailing from opposite poles — posing dolefully on the iceberg. Alshehhi has said that “it will be cheaper to bring in these icebergs” than to desalinate seawater — a common technique in the Middle East.
    World’s largest ice sheet threatened by warm water surge
    Desalination provides at least 35 trillion litres of drinking water globally every year. Birkhold notes that it is prohibitively expensive in many places. It also relies on fossil fuels for energy, and pollutes the ocean with excess salt. But he offers limited information on other, perhaps more effective, alternatives to iceberg towing, such as recycling municipal wastewater or tapping brackish water for crop irrigation. He offers no data on more esoteric sources of water, such as fog harvesting, used by remote communities in Chile, Morocco and South Africa. Nor does he address initiatives to reduce water waste or increase efficiency of use.Birkhold speculates that if iceberg harvesting succeeds, it might not remain the province of quixotic entrepreneurs — bigger, water-hungry enterprises might muscle in with their “deep coffers and profit-driven approach”. The race is largely unregulated: few national laws address iceberg use, and no international agreements clarify who can capture and sell these freshwater resources. If they are to be exploited fairly, the author concludes, “we need to decide who gets to use icebergs — and how and how many — in a way that is just and equitable”.Birkhold is engagingly honest about potential pitfalls. But if the enterprise could succeed, enormous quantities of fresh water that would otherwise melt into the ocean could be delivered to parched regions. Kudos to the author for diving in — whether or not it is ever realized.

    Competing Interests
    The author declares no competing interests. More