Symmetry-reduction enhancement of nitrate removal on record-breaking layered yttrium hydroxide adsorbents

Xinyue Sheng, Yu Wang and Jun Xu*
Tianjin Key Lab for Rare Earth Materials and Applications, School of Materials Science and Engineering & National Institute for Advanced Materials, Nankai University, Tianjin 300350, P.R. China. E-mail: junxu@nankai.edu.cn

Received 7th June 2025 , Accepted 23rd July 2025

First published on 25th July 2025


Abstract

A new class of layered metal hydroxide materials, namely layered yttrium hydroxides (LYH-X, X are anions such as Cl and Br), are reported to be excellent adsorbents to capture nitrate-nitrogen from neutral water. More importantly, the adsorption properties are correlated with the crystal symmetry of adsorbents: both the adsorption capacity and rate constant of LYH-Cl (orthorhombic P21212 space group) are almost twice of those of LYH-Br (monoclinic P21 space group). A comprehensive study combining multinuclear solid-state NMR spectroscopy and other multiscale characterization techniques was then performed to understand the origin of the symmetry-related adsorption behaviors. The results reveal an increasing trend of change in local environments of Y(OH)7·H2O, Y(OH)8·H2O, and Y(OH)8, implying that the adsorbed nitrate anions are located within the pocket constructed by alternating Y(OH)8·H2O and Y(OH)8. The splitting of the 89Y NMR peak of Y(OH)8·H2O of LYH-Cl after adsorption is consistent with the reduction of crystal symmetry from P21212 to its translationengleiche subgroup P2, providing an additional driving force for nitrate capture. This work thus not only discovers a nitrate adsorbent with a superior capacity of 44.56 ± 0.17 mg g−1, but also demonstrates a new concept of symmetry-driven adsorption enhancement.



New concepts

We propose a crystal-symmetry-driven principle to design and optimize the properties of layered rare-earth hydroxides, moving beyond conventional composition-focused strategies. We demonstrate that the symmetry reduction of LYH-Cl during nitrate adsorption enables a record-high adsorption capacity of 44.56 ± 0.17 mg g−1 at pH 7, which is more than twice that of its low-symmetry analog of LYH-Br. In addition, the LYH-Cl adsorbent exhibits a doubled adsorption rate constant compared to the LYH-Br adsorbent. The results fundamentally challenge the conventional optimization strategies for layered metal hydroxides, which have historically prioritized the dimensional matching via interlayer spacing modulation or ionic radius selection, while overlooking the effect of crystal symmetry. The symmetry-driven principle thus offers a previously unexplored dimension for searching adsorbents with exceptional potential for environmental remediation, broadening their applicability in sustainable chemistry, materials science, etc.

Introduction

The escalating prevalence of nitrogen pollutants such as nitrate (NO3) in aquatic ecosystems has emerged as a critical environmental challenge, particularly in regions with intensive agricultural activities. In these areas, groundwater nitrate concentrations often exceed 100 mg L−1, while the World Health Organization's (WHO) recommended safety limit for nitrate-nitrogen is only 11.3 mg L−1.1,2 This alarming situation highlights the urgent need for highly effective nitrate removal techniques. The widely applied techniques of biological denitrification,3–6 chemical precipitation7 and electrochemical reduction8 face limitations like secondary pollution, high energy consumption and inadequate removal, especially when dealing with the complex water matrices and high concentration pollutants. As a promising alternative, the adsorption technique has gained attention due to its simplicity, cost-effectiveness, and high efficiency. Among various nitrate adsorbents such as activated carbon,9–13 zeolites,14–17 metal–organic frameworks (MOFs),18–20 etc.,21,22 layered double hydroxides (LDHs) and their derived materials can exhibit remarkable adsorption properties (Table S1, ESI).23–26 In particular, the benchmark Cl-LDHs-C500 adsorbent possesses a nitrate-nitrogen uptake of 15.83 mg g−1 at pH 5 within 300 minutes.27 However, searching novel adsorbents with improved adsorption capacity and kinetics is still desirable for nitrate removal.

In recent years, a newer class of layered metal hydroxides, namely layered rare earth hydroxides (LREH-X, in which RE3+ are rare earth cations and X are anions),28–30 have been reported as promising adsorbents via anion exchange.31–33 More importantly, the crystal symmetry of LREH-X can vary with the anions despite similar layered structures,34 providing new opportunities to further tune the adsorption properties by selecting different X anions. Herein, we conducted a systematic study of nitrate adsorption performances on two representative layered rare earth hydroxides, namely LYH-Cl and LYH-Br, and then explored the effect of crystal symmetry on nitrate adsorption. The results obtained show that this work not only deliver adsorbents with high nitrate capture performances but also demonstrate a new concept of symmetry-driven adsorption enhancement.

Results and discussion

As shown in Fig. 1, LYH-Cl and LYH-Br share many structural and functional similarities to LDHs such as metal hydroxide layers and anion-exchange capacity, but exhibit higher intralayer heterogeneity. Distinct to the uniform octahedral M(OH)6 (M = metal ions) in LDHs, LYH-Cl and LYH-Br consist of Y3+ ions in three coordination types: eight-coordinated Y(OH)7·H2O and Y(OH)8, nine-coordinated Y(OH)8·H2O. These polyhedra share edges to construct quasi-hexagonal layers, with the coordinated water molecules oriented towards the intralayer region, which are different from the flat layers of LDHs.34 However, LYH-Cl adopts a higher crystal symmetry of P21212 than LYH-Br (P21). We have previously found that such a symmetry difference may lead to inequivalent anion-exchange behaviour, e.g., only the transformation from a high-symmetry phase (LYH-Cl) to a low-symmetry phase (LYH-Br) is favourable.34
image file: d5mh01085c-f1.tif
Fig. 1 Crystal structure of LYH-Cl and LYH-Br viewed perpendicular and parallel to the yttrium hydroxide layers, respectively. Yttrium ions are represented as black spheres, oxygen atoms as red spheres, chloride ions as bright green spheres, and bromide ions as dark green spheres. Hydrogen atoms are omitted for clarity. Color-coded polyhedra represent different crystallographically distinct yttrium coordination sites for visual clarity.

The anion-exchange characteristics of LREH-X compounds ensure them as promising candidates for nitrate removal with a high theoretical capacity of 43.01 mg g−1 (nitrogen content) for LYH-Cl and 37.85 mg g−1 for LYH-Br, respectively, if the nitrate adsorption is solely due to anion exchange.30,32,34–36 To confirm that, we first measured the nitrate adsorption performances of two materials under practical wastewater-treatment conditions (pH 7, 25 °C, 250 mg L−1 initial nitrogen concentration, 12 h) and found that both materials exhibit exceptional nitrogen uptake (26.93 mg g−1 for LYH-Cl and 16.28 mg g−1 for LYH-Br, respectively). We then explored the pH-dependence of nitrogen uptake on LYH-Cl and LYH-Br (Fig. 2a). The monotonic decrease of nitrogen uptake between pH 5 and pH 10 for both materials is consistent with the trend of change in ζ-potential.37 However, the nitrogen uptake of LYH-Br decays faster than that of LYH-Cl, especially at high pH. Despite similar ζ-potentials at pH 7 (≈ 10 mV), the uptake of LYH-Cl (44.56 ± 0.17 mg g−1) is more than twice the uptake of LYH-Br (20.87 ± 0.24 mg g−1). Kinetic studies (Fig. 2b) illustrate that the nitrate adsorption adopts a pseudo-second-order behaviour for both adsorbents, suggesting chemisorption as the rate-limiting step,38–40 and the rate constant of LYH-Cl (k2 = 0.0018 g (mg min)−1) is twice that of LYH-Br (k2 = 0.0009 g (mg min)−1). The adsorption isotherms can be fitted well with the Langmuir model (Fig. 2c), with a fitted maximum capacity (Qmax) of 45.63 mg g−1 for LYH-Cl. In contrast, LYH-Br exhibits a Qmax of 20.59 mg g−1. The agreement with the Langmuir model validates a monolayer adsorption mechanism, where nitrate anions tend to bind to homogeneous surfaces without obvious intermolecular interactions (see the Freundlich model fitting shown in Fig. S1, ESI). To eliminate the effect of possible impurities of LYH-Cl and LYH-Br, we further synthesized Y2O3 and Y(OH)3 (Fig. S2, ESI) and found the two materials exhibit very small nitrogen uptake under the same condition (< 5 mg g−1). To further evaluate the suitability of LYH-Cl and LYH-Br as nitrate adsorbents in wastewater treatment, cyclic adsorption–desorption experiments were conducted by using saturated NaCl or NaBr (25 °C, 12 h) for desorption. A decreasing trend of adsorption capacity is observed for both adsorbents (Fig. S3 and S4, ESI) due to incomplete desorption of nitrate anions and the CO32− contamination from air, which is common for anion-exchangeable layered metal hydroxides. It is worth noting that these layered metal hydroxides can be regenerated by the calcination-rehydration process, which is more effective in restoring the pristine structure and adsorption capacity. The influence of competing anions (Cl, Br, SO42−, PO43−, and HCO3) was also examined with a concentration of 100 mg L−1 for NO3–N and 200 mg L−1 for the competing anion (Fig. S5, ESI). It is striking that LYH-Cl retains more than 60% of adsorption capacity even in the presence of SO42− and PO43−, in which the high charge density of these multivalent anions yields a strong electrostatic affinity to dominate site competition.19 And the leaching amount of Y in adsorption–desorption cycles was not significant (Fig. S6, ESI).


image file: d5mh01085c-f2.tif
Fig. 2 (a) Adsorption capacity of NO3–N on LYH-Cl (left) and LYH-Br (right) as a function of pH values (2.5 g L−1 adsorbent, 500 mg L−1 NO3–N, 12 h, 25 °C). (b) NO3–N adsorption capacity of LYH-Cl (left) and LYH-Br (right) as a function of time fitted with pseudo-first-order and pseudo-second-order kinetic models (2.5 g L−1 adsorbent, 500 mg L−1 NO3–N, 12 h, 25 °C). (c) Adsorption isotherm of LYH-Cl (left) and LYH-Br (right) fitted with the Langmuir model (2.5 g L−1 adsorbent, Ce = 50–1400 mg L−1, 12 h, 25 °C; error bars are enlarged by a factor of five for clarity).

To explore the molecular mechanism for nitrate adsorption, we systematically compared the structure of adsorbents before and after the reaction by powder X-ray diffraction (PXRD), scanning electron microscopy-energy dispersive spectroscopy (SEM-EDS), X-ray photoelectron spectroscopy (XPS) and Fourier-transform infrared (FTIR) spectroscopy. As shown in Fig. 3a, most of the diffraction peaks were invisible after adsorption even at a very low nitrogen concentration of 50 mg L−1, leaving only (00n) diffraction peaks characteristic of layered structures, which can be induced by displacement between adjacent yttrium hydroxide layers. In contrast, fine features were still present in the PXRD pattern of LYH-Br after adsorption even at the highest nitrogen concentration, implying a significantly lower stacking disorder of layers. As a result, the degree of stacking disorder seems to positively correlate with the nitrate adsorption capacity. Additional PXRD data (Fig. S7, ESI) imply that the stacking disorder of LYH-Cl emerges immediately after adsorption, while such disorder can only be observed at a very long adsorption time for LYH-Br (Fig. S8, ESI). SEM images show that both adsorbents retain their initial platelet morphology after adsorption (Fig. 3b) but the surface N/Cl atomic ratio is much higher than the surface N/Br atomic ratio (EDS mapping, Fig. S9–S11, ESI). XPS analyses also confirm the significantly higher nitrogen uptake on LYH-Cl under the same conditions (Fig. 3c and Fig. S12–S14, ESI). Different from EDS and XPS, the FTIR technique can provide insights into the structural changes of bulk materials. Fig. 3d indicates that the nitrate ions have been captured by two adsorbents by the emergence of strong NO3 vibrational modes at 1384 cm−1 (ν3 asymmetric stretch) and 1084 cm−1 (ν1 symmetric stretch).36


image file: d5mh01085c-f3.tif
Fig. 3 (a) PXRD patterns of LYH-Cl (left) and LYH-Br (right) before and after nitrate adsorption (Ce = 50–1200 mg L−1, pH 7, 12 h). (b) SEM images of LYH-Cl (left) and LYH-Br (right) before and after nitrate adsorption (2.5 g L−1 adsorbent, 1200 mg L−1 NO3–N, pH 7, 12 h). (c) Cl 2p and N 1s XPS spectra of LYH-Cl (left) before and after nitrate adsorption, Br 3d and N 1s XPS spectra of LYH-Br (right) before and after nitrate adsorption (2.5 g L−1 adsorbent, 1200 mg L−1 NO3–N, pH 7, 12 h). (d) FTIR spectra of LYH-Cl (left) and LYH-Br (right) before and after nitrate adsorption as a function of time (2.5 g L−1 adsorbent, 500 mg L−1 NO3–N, pH 7).

As discussed above, routine characterization techniques such as PXRD, SEM-EDS, XPS and FTIR only provide limited insights into the structural changes after adsorption because they are not sensitive to or lack the resolution necessitated to distinguish local environments. To bridge this gap, we developed solid-state nuclear magnetic resonance (ssNMR) methodology to provide atomic-level structural information within the LYH-X materials, including site symmetry, displacement of hydroxide layers and anions, identity of hydrogen species and their spatial proximity, structural evolution during calcination, state of anions after exchange, etc.32,34,41–44 Herein, we first conducted 89Y ssNMR experiments to probe the change of the local Y3+ environment within hydroxide layers. As shown in Fig. 4a, the 89Y NMR spectrum of LYH-Cl exhibits dramatic and site-specific changes after adsorption. The 89Y peak of the Y(OH)7·H2O environment (Y1 site) is only broadened, while the peaks of Y(OH)8 (Y2 site) and Y(OH)8·H2O (Y3 site) display significant changes: the peak of the Y2 site moves to the shielded region and the peak of the Y3 site moves to the opposite (deshielded) direction, splitting into two peaks. As 89Y isotropic chemical shift (δiso(89Y)) values are very sensitive to the local environment of Y3+, decreasing with the increasing coordination number,45 the shift of the Y2 peak from 125.8 ppm to 49.6 ppm suggests that this site must strongly interact with the adsorbed nitrate ions to form a quasi-nine-coordinated environment, with δiso(89Y) similar to that of the Y(OH)9 (68 ppm) in hexagonal Y(OH)3.46 Moreover, the splitting of the Y3 peak (14.7 ppm) to two 89Y peaks with equal intensity (24.3 ppm and 29.7 ppm) unambiguously verifies the symmetry reduction of LYH-Cl during nitrate adsorption. According to crystallography, every space group can be facilely converted to its maximal subgroups by the corresponding transformation matrix:47 translationengleiche subgroup(s) (t-subgroup) whose point group is decreased, and klassengleiche subgroup(s) (k-subgroup) whose point group is retained. As Scheme S1 (ESI) illustrates, there are two t-subgroups of P21212 (#18): P21 (#4) and P2 (#3). The symmetry reduction from P21212 to P21 induces the splitting of Wyckoff position 4c to two Wyckoff positions 2a and 2a, giving rise to the splitting of the single 89Y NMR peak of Y(OH)7·H2O for LYH-Cl to two 89Y NMR peaks of Y(OH)7·H2O for LYH-Br. Similarly, the symmetry reduction from P21212 to P2 induces additional splitting of Wyckoff position 2a to two Wyckoff positions 1a and 1d (2b to 1b and 1c), consistent with the splitting of the single 89Y NMR peak of Y(OH)8·H2O (2b) to two 89Y NMR peaks (1b and 1c) for LYH-Cl after adsorption. We should point out that the splitting of other 89Y NMR peaks is not observed due to small δiso(89Y) differences and/or significant peak broadening. In contrast, the 89Y NMR spectrum of LYH-Br is almost not altered by nitrate adsorption except the peak broadening and the emergence of a weak and very broad peak centered at 66.7 ppm. Based on the relative intensity, it seems that some Y(OH)8 spheres (Y3 site) are converted to quasi-nine-coordinated spheres with a δiso(89Y) of 66.7 ppm, whereas the structural changes on other sites are negligible. We further monitored the replacement of anions by 35Cl (Fig. 4b) and 79Br (Fig. 4c) ssNMR experiments. The 35Cl NMR signal decays dramatically after adsorption and becomes almost invisible at higher nitrogen concentration for LYH-Cl, whereas the 79Br NMR signal persists even at very high nitrogen concentrations. Such observation is consistent with the adsorption capacity of two materials: all Cl anions can be substituted by NO3 in LYH-Cl and only half of the Br anions can be substituted by NO3 in LYH-Br.


image file: d5mh01085c-f4.tif
Fig. 4 (a) 89Y NMR spectra of LYH-Cl (left) and LYH-Br (right) before and after nitrate adsorption (Ce = 50–1200 mg L−1, 2.5 g L−1 adsorbent, pH 7, 12 h). (b) 35Cl NMR spectra of LYH-Cl (*: NaCl) before and after nitrate adsorption (Ce = 50–1200 mg L−1, 2.5 g L−1 adsorbent, pH 7, 12 h). (c) 79Br NMR spectra of LYH-Br (*: NaBr) before and after nitrate adsorption (Ce = 50–1200 mg L−1, 2.5 g L−1 adsorbent, pH 7, 12 h).

The molecular mechanism of nitrate adsorption on two LYH-X compounds can now be proposed as follows (Fig. 5): the NO3 anions insert into the interlayer space and replace existing Cl or Br anions during adsorption. The adsorbed nitrate anions reside within the pockets constructed by alternating Y(OH)8 and Y(OH)8·H2O polyhedra, with weak coordination on the Y3+ ion of Y(OH)8 and hydrogen bonding to coordinated water molecules of Y(OH)8·H2O. However, the crystal symmetry of LYH-Cl is reduced by the displacement between neighbouring layers accompanying with nitrate insertion. In addition, the increase of entropy of the LYH-Cl adsorbent by lowering its crystal symmetry from P21212 to P2 can provide extra driving force for nitrate adsorption. This directional transition follows Bärnighausen's group-subgroup hierarchy described previously in the literature,48 where asymmetric energy barriers are present between the structural transformation between a space group and its t-subgroup. Both factors contribute to the higher adsorption capacity and faster adsorption kinetics of LYH-Cl than LYH-Br.


image file: d5mh01085c-f5.tif
Fig. 5 Schematic illustration of the nitrate adsorption mechanism of LYH-Cl (left) and LYH-Br (right).

Conclusions

This comprehensive study reports a novel layered metal hydroxide adsorbent (LYH-Cl) with a record-breaking capacity of 44.56 ± 0.17 mg g−1 at pH 7. By comparing LYH-Cl (orthorhombic P21212 space group) with its lower-symmetry analog LYH-Br (monoclinic P21 space group), it is found that the reduction of crystal symmetry during the anion-exchange reaction is responsible for the doubled capacity and rate constant of LYH-Cl, unraveled by a multiscale investigation integrating PXRD, SEM-EDS, XPS, FTIR, and especially ssNMR results. This work not only develops high-performance materials for environmental remediation, but also provides an additional symmetry-driven principle to design and optimize the properties of layered materials, opening avenues to broaden their applicability in sustainable chemistry, materials science, etc. And we will continue to work on the scalability and optimization of structuring/regeneration conditions of the LYH-Cl adsorbent to check its feasibility for large-scale applications.

Author contributions

Conceptualization: JX; methodology: XS; supervision: JX; investigation: XS, YW; formal analysis: XS; resources: JX; writing – original draft: XS; writing – review and editing: XS, JX; project administration: JX; and funding acquisition: JX.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the ESI.

Acknowledgements

We thank the financial support from the National Natural Science Foundation of China (grant no. 22471128).

References

  1. J. N. Galloway, A. R. Townsend, J. W. Erisman, M. Bekunda, Z. Cai, J. R. Freney, L. A. Martinelli, S. P. Seitzinger and M. A. Sutton, Transformation of the Nitrogen Cycle: Recent Trends, Questions, and Potential Solutions, Science, 2008, 320(5878), 889–892 CrossRef.
  2. World Health, O. Nitrate and nitrite in drinking-water: background document for development of WHO guidelines for drinking-water quality; World Health Organization: Geneva, 2003, 2003.
  3. H. Daims, E. V. Lebedeva, P. Pjevac, P. Han, C. Herbold, M. Albertsen, N. Jehmlich, M. Palatinszky, J. Vierheilig, A. Bulaev, R. H. Kirkegaard, M. von Bergen, T. Rattei, B. Bendinger, P. H. Nielsen and M. Wagner, Complete nitrification by Nitrospira bacteria, Nature, 2015, 528(7583), 504–509 CrossRef PubMed.
  4. C. E. Lawson, S. Wu, A. S. Bhattacharjee, J. J. Hamilton, K. D. McMahon, R. Goel and D. R. Noguera, Metabolic network analysis reveals microbial community interactions in anammox granules, Nat. Commun., 2017, 8(1), 15416 CrossRef.
  5. C. Hellinga, A. A. J. C. Schellen, J. W. Mulder, M. C. M. van Loosdrecht and J. J. Heijnen, The sharon process: an innovative method for nitrogen removal from ammonium-rich waste water, Water Sci. Technol., 1998, 37(9), 135–140 CrossRef.
  6. J. Lugo-Arias, J. González-Álvarez, A. Maturana, J. Villa-Parejo and C. Barraza-Heras, Removal of nitrate and phosphate from aqueous solutions using bioadsorbents derived from agro-industrial wastes of rice husk and corn stalk, Biomass Convers. Biorefin., 2025, 15(12), 19453–19475 CrossRef.
  7. M. Bahrami and M. J. Amiri, Nitrate removal from contaminated waters using modified rice husk ash by Hexadecyltrimethylammonium bromide surfactant, React. Kinet., Mech. Catal., 2022, 135(1), 459–478 CrossRef CAS.
  8. S. Garcia-Segura, M. Lanzarini-Lopes, K. Hristovski and P. Westerhoff, Electrocatalytic reduction of nitrate: Fundamentals to full-scale water treatment applications, Appl. Catal., B, 2018, 236, 546–568 CrossRef.
  9. M. Lebrun, S. Renouard, D. Morabito and S. Bourgerie, Sorption capacities of various activated carbons towards nitrates: effects of nitrate concentration, pH, time and co-existing ions, Int. J. Environ. Sci. Technol., 2023, 20(12), 13033–13044 CrossRef.
  10. A. R. Satayeva, C. A. Howell, A. V. Korobeinyk, J. Jandosov, V. J. Inglezakis, Z. A. Mansurov and S. V. Mikhalovsky, Investigation of rice husk derived activated carbon for removal of nitrate contamination from water, Sci. Total Environ., 2018, 630, 1237–1245 CrossRef PubMed.
  11. M. J. Ahmed, B. H. Hameed and M. A. Khan, Recent advances on activated carbon-based materials for nitrate adsorption: A review, J. Anal. Appl. Pyrolysis, 2023, 169, 105856 CrossRef.
  12. P. Mehrabinia and E. Ghanbari-Adivi, Examining nitrate surface absorption method from polluted water using activated carbon of agricultural wastes, Model. Earth Syst. Environ., 2022, 8(2), 1553–1561 CrossRef.
  13. C. Yuan, F. S. Cannon and Z. Zhao, Removing nitrate with coconut activated carbon, tailored with quaternary ammonium epoxide compounds: Effect of base or acid carbon pretreatment, J. Environ. Manage., 2019, 234, 21–27 CrossRef PubMed.
  14. S. Wang and Y. Peng, Natural zeolites as effective adsorbents in water and wastewater treatment, Chem. Eng. J., 2010, 156(1), 11–24 CrossRef.
  15. A. F. Florea, C. Lu and H. C. B. Hansen, A zero-valent iron and zeolite filter for nitrate recycling from agricultural drainage water, Chemosphere, 2022, 287, 131993 CrossRef.
  16. J. M. Mir, M. A. Mir, S. A. Majid and B. A. Malik, Solid state electron density aspects of mordenite with experimentally evaluated nitrate and phosphate sorption, Mater. Sci. Eng. B, 2021, 263, 114865 CrossRef.
  17. P. Kuang, N. Chen, C. Feng, M. Li, S. Dong, L. Lv, J. Zhang, Z. Hu and Y. Deng, Construction and optimization of an iron particle–zeolite packing electrochemical–adsorption system for the simultaneous removal of nitrate and by-products, J. Taiwan Inst. Chem. Eng., 2018, 86, 101–112 CrossRef.
  18. P. Liu, J. Yan, H. Huang and W. Song, Cu/Co bimetallic conductive MOFs: Electronic modulation for enhanced nitrate reduction to ammonia, Chem. Eng. J., 2023, 466, 143134 CrossRef.
  19. I. Aswin Kumar, A. Jeyaseelan, N. Viswanathan, M. Naushad and A. J. M. Valente, Fabrication of lanthanum linked trimesic acid as porous metal organic frameworks for effective nitrate and phosphate adsorption, J. Solid State Chem., 2021, 302, 122446 CrossRef.
  20. I. A. Kumar, A. Jeyaseelan, S. Ansar and N. Viswanathan, A facile synthesis of 2D iron bridged trimesic acid based MOFs for superior nitrate and phosphate retention, J. Environ. Chem. Eng., 2022, 10(2), 107233 CrossRef CAS.
  21. D. Mohan, A. Sarswat, Y. S. Ok and C. U. Pittman, Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent – A critical review, Bioresour. Technol., 2014, 160, 191–202 CrossRef CAS.
  22. C. Guo, L. Shen, Y. Tang, M. Chen, X. Zhou, F. Ciucci and Z. Tang, Ru single atoms anchored on Co-NC for nitrate electroreduction toward NH3 synthesis, Nano Res., 2025 DOI:10.26599/NR.2025.94907623.
  23. S. Li, X. Ma, Z. Ma, X. Dong, Z. Wei, X. Liu and L. Zhu, Mg/Al-layered double hydroxide modified biochar for simultaneous removal phosphate and nitrate from aqueous solution, Environ. Technol. Innovation, 2021, 23, 101771 CrossRef CAS.
  24. K.-H. Goh, T.-T. Lim and Z. Dong, Application of layered double hydroxides for removal of oxyanions: A review, Water Res., 2008, 42(6), 1343–1368 CrossRef CAS PubMed.
  25. S. Lu, Q. Zhu and R. Li, Selective adsorption of nitrate in water by organosilicon quaternary ammonium salt modified derived nickel-iron layered double hydroxide: Adsorption characteristics and mechanism, J. Colloid Interface Sci., 2023, 652, 1481–1493 CrossRef CAS PubMed.
  26. M. Elhachemi, Z. Chemat-Djenni, D. Chebli, A. Bouguettoucha and A. Amrane, Synthesis and physicochemical characterization of new calcined layered double hydroxide Mg Zn Co Al-CO3; classical modeling and statistical physics of nitrate adsorption, Inorg. Chem. Commun., 2022, 145, 109549 CrossRef CAS.
  27. S. Mariska, Z. Jin-Wei, H. H. Chien, D. M. Ngoc, N. D. Hai and H.-P. Chao, Adsorptive removal of phosphate and nitrate by layered double hydroxides through the memory effect and in situ synthesis, Appl. Water Sci., 2025, 15(2), 42 CrossRef CAS.
  28. X. Wang, M. S. Molokeev, Q. Zhu and J.-G. Li, Controlled Hydrothermal Crystallization of Anhydrous Ln2(OH)4SO4 (Ln = Eu–Lu, Y) as a New Family of Layered Rare Earth Metal Hydroxides, Chem. – Eur. J., 2017, 23(63), 16034–16043 CrossRef CAS PubMed.
  29. K.-H. Lee and S.-H. Byeon, Synthesis and Aqueous Colloidal Solutions of RE2(OH)5NO3·nH2O (RE = Nd and La), Eur. J. Inorg. Chem., 2009, 4727–4732 CrossRef CAS.
  30. F. Gándara, J. Perles, N. Snejko, M. Iglesias, B. Gómez-Lor, E. Gutiérrez-Puebla and M. Á. Monge, Layered Rare-Earth Hydroxides: A Class of Pillared Crystalline Compounds for Intercalation Chemistry, Angew. Chem., Int. Ed., 2006, 45(47), 7998–8001 CrossRef.
  31. F. Chen, Y. Liang, Z. Zhou, T. Yu, Y. Mei, Z. Zhou, B. Chen, Y. Liang and Y. Wang, Monolayer calcium-doped layered yttrium hydroxide adsorbent for rapid and efficient removal of fluoride from industrial wastewater, Chem. Eng. Sci., 2024, 300, 120630 CrossRef CAS.
  32. K. Wu and J. Xu, The Ordering and Arrangement of Intercalated Carboxylate Ions in Anion-Exchangeable Layered Yttrium Hydroxides, Eur. J. Inorg. Chem., 2024, e202400477 CrossRef CAS.
  33. M. Kim, H. Kim and S.-H. Byeon, Layered Yttrium Hydroxide l-Y(OH)3 Luminescent Adsorbent for Detection and Recovery of Phosphate from Water over a Wide pH Range, ACS Appl. Mater. Interfaces, 2017, 9(46), 40461–40470 CrossRef CAS.
  34. Z. Feng, D. Xiao, Z. Liu, G. Hou and J. Xu, “X Factor” in the Structure and Anion Exchange of Layered Yttrium Hydroxides, J. Phys. Chem. C, 2021, 125(13), 7251–7258 CrossRef CAS.
  35. B. Li, L. Zheng, Y. Gong and P. Kang, Intercalation-driven tunability in two-dimensional layered materials: Synthesis, properties, and applications, Mater. Today, 2024, 81, 118–136 CrossRef CAS.
  36. L. J. McIntyre, L. K. Jackson and A. M. Fogg, Ln2(OH)5NO3·xH2O (Ln = Y, Gd−Lu): A Novel Family of Anion Exchange Intercalation Hosts, Chem. Mater., 2008, 20(1), 335–340 CrossRef.
  37. A. Bhatnagar and M. Sillanpää, A review of emerging adsorbents for nitrate removal from water, Chem. Eng. J., 2011, 168(2), 493–504 CrossRef.
  38. Y. S. Ho and G. McKay, Pseudo-second order model for sorption processes, Process Biochem., 1999, 34(5), 451–465 CrossRef.
  39. S. Azizian, Kinetic models of sorption: a theoretical analysis, J. Colloid Interface Sci., 2004, 276(1), 47–52 CrossRef PubMed.
  40. N. Ayawei, A. N. Ebelegi and D. Wankasi, Modelling and Interpretation of Adsorption Isotherms, J. Chem., 2017, 2017(1), 3039817 Search PubMed.
  41. Y. Liu, X. Sheng, H. Ding and J. Xu, A comprehensive solid-state NMR and theoretical modeling study to reveal the structural evolution of layered yttrium hydroxide upon calcination, J. Magn. Reson. Open, 2024, 20, 100155 CrossRef.
  42. Y. Liu, X. Sheng, K. Zhu, X. Li, X. Yao, G. Hou and J. Xu, Mapping Hydrogens in Hydrous Materials, J. Phys. Chem. C, 2023, 127(18), 8640–8648 CrossRef.
  43. Y. Liu, S. Jiang and J. Xu, Exploring the intercalation chemistry of layered yttrium hydroxides by 13C solid-state NMR spectroscopy, Magn. Reson. Lett., 2022, 2(3), 186–194 CrossRef.
  44. Z. Liu, L. Liang, D. Xiao, Y. Ji, Z. Zhao, J. Xu and G. Hou, 89Y chemical shift anisotropy: a sensitive structural probe of layered yttrium hydroxides revealed by solid-state NMR spectroscopy and DFT calculations, Phys. Chem. Chem. Phys., 2021, 23(48), 27244–27252 RSC.
  45. J. Xu, S. Jiang and Y. Du, Unravelling the Mystery of Solid Solutions: A Case Study of 89Y Solid-State NMR Spectroscopy, ChemPhysChem, 2020, 21(9), 825–836 CrossRef CAS PubMed.
  46. C. V. Chandran, H. Schreuders, B. Dam, J. W. G. Janssen, J. Bart, A. P. M. Kentgens and P. J. M. van Bentum, Solid-State NMR Studies of the Photochromic Effects of Thin Films of Oxygen-Containing Yttrium Hydride, J. Phys. Chem. C, 2014, 118(40), 22935–22942 CrossRef CAS.
  47. M. I. Aroyo, J. M. Perez-Mato, C. Capillas, E. Kroumova, S. Ivantchev, G. Madariaga, A. Kirov and H. Wondratschek, Z. Kristallogr., 2006, 221(1), 15–27 CAS.
  48. H. Bärnighausen, Group-subgroup relations between space groups: a useful tool in crystal chemistry, Match-Commun. Math. Comput. Chem., 1980, 9, 139–175 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5mh01085c

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.